Next Article in Journal
The Role of Echocardiography in the Diagnosis of Cardiac Involvement in a Rare Systemic Condition: The Carcinoid Heart Disease: A Case Report and Review of Literature
Next Article in Special Issue
The Genetic and Epigenetic Footprint in Idiopathic Pulmonary Fibrosis and Familial Pulmonary Fibrosis: A State-of-the-Art Review
Previous Article in Journal
Pediatric Low-Grade Glioma Surgery with Sodium Fluorescein: Efficient Localization for Removal and Association with Intraoperative Pathological Sampling
Previous Article in Special Issue
Pulmonary Hypertension Associated Genetic Variants in Sarcoidosis Associated Pulmonary Hypertension
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Genetics in Idiopathic Pulmonary Fibrosis: A Clinical Perspective

1
2nd Pulmonary Medicine Department, General University Hospital “Attikon”, Medical School, National and Kapodistrian University of Athens, 12462 Athens, Greece
2
Département de Génétique, APHP Hôpital Bichat, Université de Paris, 75018 Paris, France
3
INSERM UMR 1152, Université de Paris, 75018 Paris, France
4
Service de Pneumologie A, INSERM UMR_1152, Centre de Référence des Maladies Pulmonaires Rares, FHU APOLLO, APHP Hôpital Bichat, Sorbonne Université, 75018 Paris, France
5
7th Pulmonary Department, Athens Chest Hospital “Sotiria”, 11527 Athens, Greece
6
Peditric Pulmonology Department and Reference Centre for Rare Lung Diseases RespiRare, INSERM UMR_S933 Laboratory of Childhood Genetic Diseases, Armand Trousseau Hospital, Sorbonne University and APHP, 75012 Paris, France
7
Paediatrics and Paediatric Respirology, Imperial College, Imperial Centre for Paediatrics and Child Health, Royal Brompton Harefield NHS Foundation Trust, London SW3 6NP, UK
8
Department of Pediatric Pneumology, Dr von Hauner Children’s Hospital, Ludwig-Maximilians-University, German Center for Lung Research, 80337 Munich, Germany
9
Department of Rheumatology, INSERM U1152, APHP Hôpital Bichat-Claude Bernard, Université de Paris, 75018 Paris, France
*
Author to whom correspondence should be addressed.
Diagnostics 2022, 12(12), 2928; https://doi.org/10.3390/diagnostics12122928
Submission received: 14 October 2022 / Revised: 17 November 2022 / Accepted: 18 November 2022 / Published: 23 November 2022
(This article belongs to the Special Issue Molecular Diagnosis of Interstitial Lung Disease)

Abstract

:
Background: Unraveling the genetic background in a significant proportion of patients with both sporadic and familial IPF provided new insights into the pathogenic pathways of pulmonary fibrosis. Aim: The aim of the present study is to overview the clinical significance of genetics in IPF. Perspective: It is fascinating to realize the so-far underestimated but dynamically increasing impact that genetics has on aspects related to the pathophysiology, accurate and early diagnosis, and treatment and prevention of this devastating disease. Genetics in IPF have contributed as no other in unchaining the disease from the dogma of a “a sporadic entity of the elderly, limited to the lungs” and allowed all scientists, but mostly clinicians, all over the world to consider its many aspects and “faces” in all age groups, including its co-existence with several extra pulmonary conditions from cutaneous albinism to bone-marrow and liver failure. Conclusion: By providing additional evidence for unsuspected characteristics such as immunodeficiency, impaired mucus, and surfactant and telomere maintenance that very often co-exist through the interaction of common and rare genetic variants in the same patient, genetics have created a generous and pluralistic yet unifying platform that could lead to the understanding of the injurious and pro-fibrotic effects of many seemingly unrelated extrinsic and intrinsic offending factors. The same platform constantly instructs us about our limitations as well as about the heritability, the knowledge and the wisdom that is still missing.

1. Introduction

Fibrotic interstitial lung diseases (f-ILDs) are a heterogeneous group of diffuse parenchymal lung disorders that may share common overlapping clinical characteristics and, under not yet clearly identified circumstances, progression to pulmonary fibrosis [1,2,3,4]. Variable susceptibility and progression to diffuse lung fibrosis constitute the major determinant of their unpredictable clinical course and prognosis [5,6,7,8]. A major achievement of the last decades in the elucidation of their etiopathogenetic mechanisms, besides the role of environmental, aging-related, immune or other factors, is that genomics play an important, previously unsuspected, role in their development [9,10,11]. f-ILDs include idiopathic interstitial pneumonias (IIPs), autoimmune-rheumatic-disorder-related lung fibrosis, hypersensitivity pneumonia, sarcoidosis, and unclassified and rare f-ILDs [12,13,14]. Idiopathic pulmonary fibrosis (IPF), the “core” of all f-ILDs, is the most common and clinically severe, bears the histology of usual interstitial pneumonia (UIP), and although it invariably presents an ominous prognosis despite current pharmacological treatment, its clinical course is highly unpredictable, lasting at diagnosis from months to almost a decade [1,15,16]. Most of the progress on genomics in f-ILDs is in IPF, where genetic variants underlie 30% of patients at risk of developing sporadic or familial IPF (two or more cases of IPF in the same family within three degrees of relationship) [17]. The same genetic susceptibility does not lead to identical clinical, imaging, or histological phenotypes in either the sporadic or familial context, delineating the role of epigenetic factors [18,19].
The traditional view that IPF is a sporadic disease of unknown etiology limited to the lungs and almost always affecting the elderly has been challenged in the last decade by the clinical observation that it develops 10 times more often in members of families and that it may develop as part of a multisystem, inherited disease of telomeres, named telomeropathy (short telomeres syndrome, STS) where members of the same family may also present skin, mucosa, and hair abnormalities, T-cell immunodeficiency, and bone-marrow and/or liver disease [20,21,22,23,24,25]. All the above, in combination with the fact that certain forms of interstitial lung disease develop at a very early age (children’s interstitial lung disease-chILD) may be attributed to genetic predisposition, provided evidence of an inheritable component in IPF [26].Thanks to progress in genomics, it has been shown that heritability in adult IPF relates, in most cases but not all, to pathogenic variations in surfactant-related genes (SRGs) and telomere-related genes (TRG) in monogenic inheritance and a single-nucleotide polymorphism (SNP) in the promoter of the MUC5B gene encoding Mucin 5B (rs35705950 T risk allele) in polygenic inheritance (the most common genetic variant) [27]. Genetic studies in IPF and also in other f-ILDs have identified several SNPs associated with fibrosis, the rare variants conferring a higher risk while the common ones a lower risk [28]. Genome-wide-association studies continue to identify new candidate chromosomal loci to be associated with increased susceptibility to IPF [29,30,31]. The discovery in IPF that multiple genetic components, including the rarest variants, may be encountered in the same patient broadens our understanding of clinical heterogeneity [32,33,34], suggesting co-responsibility in the pathogenetic mechanisms of the disease instead of a mere cohabitation and necessitating a unifying pathogenetic theory regarding genetic risk. Differences in genomics in IPF patients may drive differences in the disease course (phenotypic unpredictability) through the diversity of pathogenetic mechanisms and in the near future may lead to a personalized approach towards precision in the cure of a currently incurable disease [32,35,36]. The aim of the study is to provide an overview of genetics in IPF from a clinical perspective.

2. Overview of IPF Pathogenesis

Despite considerable progress and decades of research unraveling the complexity of mechanisms underlying the pathophysiology of IPF, a unifying and credible hypothesis linking the multiplicity of cell types, molecular factors, and signaling pathways in its pathogenesis still remains elusive [37,38,39]. Understanding its pathology, in essence the process of lung wound healing, appears clearly the first step in the elucidation of IPF pathogenesis [40,41,42,43,44]. Parenchymal lung fibrosis, the end of the pathogenetic process, begins or appears early on in the subpleural basal and posterior peripheral areas of the lungs, sites of increased traction forces, and is characterized by extreme remodeling of the alveolar spaces and walls including the distal bronchioles (traction bronchiolectasis-bronchiectasis) [45,46,47]. The histopathological hallmarks include the bronchiolization of the distal airspaces, honeycombing, fibroblastic foci, and abnormal epithelial hyperplasia defining the UIP histology [48]. Fibrosis constitutes of an aberrant non-resolving extracellular matrix deposition and necessitates multiple cells, fibrogenic molecules, and signaling pathways triggered by several asymmetrical injuring factors, acting repeatedly on alveolar epithelial cells (AEC), according to the current pathogenetic theories [49]. Injured AECs due to exposure to inhaled adverse factors acting on a susceptible individu-al background due to ageing and susceptible genome leads to senescence of AECs and through a multitude of pathways acting on fibroblasts and myofibroblasts alters extracel-lular matrix leading to aberrant and self-perpetuating wound healing [50,51,52]. Indeed, it is known that AEC2s are progenitor cells for AEC1s (which cover most part of the alveolar surface) in a normal lung. After damage, either acute or subacute, of the alveolar epithelium, AEC2s proliferate and transdifferentiate into AEC1s. Senescent and dysfunctional AEC2s are not able to transdifferentiate into AEC1s, but activate differential pathways, e.g., those involved in epithelial–mesenchymal transition that eventually lead to aberrant repair (ECM deposition) without regeneration and to fibrosis. [53]. Genetic predisposing factors may stand alone or even coexist with most of the above predisposing conditions and through a diversity of pathogenetic mechanisms may lead to a “homogeneous” pattern of fibrotic lung disease, determine its earlier appearance, influence clinical course and prognosis, and possibly in the future direct early therapeutic measures [29,54,55,56,57]. The identification of high-risk populations may also guide early referral and the adoption of preventive measures that could postpone the development of the disease [58,59,60,61,62,63,64]. However, in familial PF, despite considerable progress, the identification of predisposing genetic variants still eludes us in a significant proportion of patients [65].

3. Historical Perspective on Genetics in IPF

Genetics have revolutionized the way we approach human disease and IPF is not an exception to this rule (Figure 1) [66,67,68]. Inheritable PF has been the focus of intense research since 1950, when the first cases of familial PF in identical twins were reported, raising questions about the presence of a “fibrosis gene” (Figure 2) [69,70].

4. Surfactant-Protein-Related Genes

In 2001, a breakthrough occurred when a mutation in the gene encoding surfactant protein C (SFTPC) in a young mother and a child with familial ILD was identified [71]. In 2002, the SFTPC mutation was identified in a large family, including adults with UIP and children with nonspecific ILD [72]. Thereafter, many studies confirmed that SFTPC mutations may cause familial lung disease in an autosomal dominant pattern with very high penetrance in all age groups, from the newborns to the elderly [73,74]. In 2009 and 2016, it was shown that very rare mutations in the genes encoding the surfactant proteins A2 and A1—respectively, SFTPA2 and SFTPA1—segregated with an ILD and adenocarcinoma phenotype in large families with multiple members with ILD occurring at a wide age range and/or adenocarcinoma only in adults [75,76]. The spectrum of SRGs implicated in inheritable PF was further expanded by the observation that bi-allelic mutations in the gene of ATP-binding cassette subfamily A, member 3 (ABCA3) protein, which plays a cardinal role in surfactant homeostasis, are associated with neonatal respiratory distress syndrome (RDS) and in “chILD” surviving into adulthood [77,78,79,80]. Mutations in NKX2-1, which encodes thyroid transcription factor-1, regulating the transcription of surfactant proteins and ABCA3, as well as many other lung proteins, was found to be associated with ILD in the context of “brain-thyroid-lung” syndrome (neurological symptoms, peripheral hypothyroidism, and f-ILD). Lung involvement ranges from neonatal RDS to adult PF inherited through an autosomal dominant pattern with variable penetrance. In 25% of patients, it develops as an isolated phenotype and in another 19% in association only with neurological symptoms [81,82,83,84]. Disease-causing mutations in the gene encoding surfactant protein D (SFTPD) have not yet been identified, whereas bi-allelic mutations in the gene encoding surfactant protein B (SFTPB) are typically associated with neonatal RDS; no adults have been reported so far [85,86]. SRG pathogenic variations seem to be implicated in the development of pulmonary fibrosis by disrupting AEC2 homeostasis. Very recently it has been shown that mutations in SFTPC lead to accumulation of the mutated protein at the plasma membrane through abnormal recycling from endosomes and from impaired internalization into multivesicular bodies with still unknown effects on the signaling within the AEC2 or its neighbors and the surface expression of otherwise unrelated proteins. Mutations in SFTPA2 and SFTPA1 are thought to abolish the secretion of mutant proteins, leading to their sequestration in the AEC2. Conversely, mutations in ABCA3 and NKX2-1 result in dysfunctional lamellar bodies and disrupt surfactant homeostasis in AEC2s, which play the cardinal role in alveolar maintenance and repair of recurrent epithelial cell injury of any etiology [87,88,89,90,91,92,93,94,95,96,97,98,99].

5. Clinical Implications of Carriership of SRG Mutations

Depending on the cohorts, the contribution of SRG mutations in monogenic pulmonary fibrosis ranges from 2 to 5%, with the exception of some populations where it may reach 25% in chILD [10,26,74,100,101,102,103,104]. Patients who are carriers of SRGs present a great heterogeneity of phenotypes, from lethal neonatal RDS to, more rarely, chILD and adult ILD [74]. In adults, a higher degree of suspicion is required, especially in early-onset disease and in familial PF, where lung fibrosis and/or pulmonary adenocarcinoma segregate in many family members. Gender and environmental exposures do not seem to play a significant role in the development of the disease. In adult patients, the age at which SRGs-ILD has been reported ranges from 19 to 71 years and depends on the underlying mutation. The median age for SFTPC is 37 years, whereas for SFTPA1/SFTPA2 is 48 years [99]. ABCA3 mutations-related ILDs have been described in cases of chILD surviving into adulthood and in adults [79,80,105,106,107]. The radiological imaging patterns conform more often to unclassifiable ILD with reticular abnormalities, ground glass opacities and scattered cystic lesions and may lead to the decision for lung biopsy, often disclosing a UIP pattern and in some association with non-specific or desquamative interstitial pneumonitis pattern or even pulmonary alveolar proteinosis (PAP)-like histology [108,109,110]. The major clinical implication so far described in SRG mutations relates to the high frequency (37%) of lung carcinoma (mostly of the adenocarcinoma type) alone (12%) or in combination with pulmonary fibrosis (25%) in SFTPA1/SFTPA2 adult carriers. Therefore patients should be regularly screened, especially after the age of 40 years [74,76,99] and family members genetically tested for the presence of SFTPA1/SFTPA2 mutations. No effective treatment exists for surfactant-related genetic disease and there is no consensus on it; studies report treatment with azithromycin or hydroxychloroquine, while others opt for the immunosuppressants azathioprine, methylprednisolone, and cyclophosphamide with no curative or long-term stabilizing effects [111,112]. Gene-based therapies as well as functional rescue therapies of misfolding ABCA3 mutations by molecular correctors are promising for the future [113,114]. For the moment, bilateral lung transplantation remains the best option for end-stage fibrotic chILD and adult ILD related to SRG mutations [74,115].

6. Telomere-Related Genes and Telomeropathy

In 2007, two independent research groups reported the association of mono-allelic telomerase mutations and the development of familial PF in adults [116,117]. Their observations initially regarded only mutations in telomerase reverse transcriptase (TERT) and the RNA component of the telomerase complex (TERC), both responsible for telomere length maintenance, and were partly inspired by the study of the genetic background of the inheritable multisystem disease dyskeratosis congenita, where mutations in the telomerase pathway were initially recognized to cause STS characterized by bone-marrow and liver failure and fibrotic ILD at a young age [118]. Dyskeratosis congenita (DC), was the first described telomeropathy; it consists of an inherited bone-marrow failure syndrome clinically characterized by a muco-cutaneous triad (skin pigmentation, nail dystrophy, mucosal leukoplakia), bone-marrow failure, immune deficiency, pulmonary fibrosis, liver cirrhosis, and the insurgence of malignancies. The discovery in the last 15 years that telomeropathy is a disease related to defects in telomere maintenance with short telomeres and mutations affecting telomerase activity, assembly, and telomere integrity has enabled scientists to consider the telomerase-complex pathway in the pathogenesis of familial PF [23,119,120,121,122,123,124,125]. The genes implicated in telomeropathy are: ACD, CTC1,DCLRE1B, DKC1, NHP2, NOP10, PARN, POT1, RPA1, RTEL1, STN1, TERC, TERT, TINF2, WRAP53, and ZCCHC8. Mono-allelic or bi-allelic pathogenic variants of a few genes are identified in patients with different clinical spectrum. Severe forms of DC and Höyeraal–Hreidarsson, Revesz, or Coats plus syndromes are more often associated with biallelic variants, while heterozygous TERT, TERC, RTEL1, PARN, ACD, NHP2, and NOP10 variants are found in PF patients(Table 1) [126]. Thereafter, the discovery of mutations in numerous genes of the telomere-maintenance pathway established the role of this pathway in the development of accelerated aging and monogenic PF and led to the acknowledgement of TRG pathogenic mutations as the major monogenic cause of IPF, the genetic basis of 30% of cases of familial PF [127,128,129,130,131,132,133,134,135,136,137,138,139]. The net effect of telomere-shortening on somatic cells beyond a critical point is DNA damage leading to apoptosis in high-turnover tissue cells and senescence in low-turnover cells such AECs, and loss of their capacity for resilience and repair leading to repeated injury due to internal or external hits and to lung parenchyma pro-fibrotic remodeling [140]. The effect of TRG germline mutations on telomere length is systemic, affecting not only the lung but also the bone marrow and liver. More than 70% of patients that carry TRG mutations develop monogenic PF, considered to be the most prevalent “telomeropathy” manifestation. Extra-pulmonary signs or diseases could also be present with variable penetrance: early hair greying, bone-marrow failure syndrome, immunodeficiency-related risk of infections, and liver disease. Interestingly, 15% of patients with familial PF are found to have short telomeres and develop extrapulmonary manifestations (personal or familial) of STS without any identifiable TRG mutation [141,142,143]. Short telomeres in familial PF patients and their families may represent a surrogate for TRG pathogenic mutations [144,145]. Shortened telomeres are also observed in a significant proportion of sporadic IPF cases, but in that case TRG pathogenetic mutations are disclosed in less than 2% of adult patients [146,147,148,149,150,151,152,153]. TRG-mutation-associated ILDs are very rare in pediatric patients outside the context of syndromes such as DC, but are increasingly recognized [154].

7. Clinical Implications of Carriership of Pathogenic Variations in Telomere-Related Genes

Approximately 30% of patients with familial PF are carriers of pathogenic variations in TRGs. Elaborate genealogical research studies have shown that a latency period of >300 years may pass before the cumulative effect of telomere shortening eventually leads to familial PF in TRG mutation carriers [141,155]. In addition to personal/familial extra-pulmonary abnormalities suggestive of STS involving the skin, the liver, and the bone marrow, the disease is uniformly progressive, irrespective of its phenotypic heterogeneity, and bears a reduced transplant-free survival time [18,19,32,156,157]. The increased risk for complications and high mortality post-lung-transplantation is due mostly to sepsis, greater than expected bone-marrow suppression and increased rates of chronic allograft dysfunction and airway complications (dehiscence and stenosis) [156,157,158,159,160]. Subclinical bone marrow and liver abnormalities, including significant T-cell immunodeficiency silently existing in these patients, may aggravate post-lung transplantation-immunosuppression-drug toxicities [161]. Therefore, telomeropathy is now recognized in addition to age, frailty, pulmonary hypertension, cardiovascular risk, and lung cancer as a parameter for which careful consideration must be given to pre-operative optimization, surgical technique, pulmonary rehabilitation, and a tailored immunosuppressive treatment protocol to produce the best post-transplantation outcomes [127,143,162,163,164,165,166]. Bone-marrow failure presenting as myelodysplastic syndrome (MDS) or acute myeloid leukemia (AML) may arise at any timepoint in the disease’s course in almost 10% of patients, especially those younger than 65 years old [18,19,32,141,167,168,169,170]. Naturally occurring somatic mutations in telomere maintenance genes observed in some patients with familial-PF potentially might rescue premature age-related clonal hematopoiesis and transformation to MDS/AML [139,171]. Solid tumors are rare; their risk appears higher in male DKC1 mutation carriers [141]. Subclinical forms of liver involvement and cryptogenic liver cirrhosis may also affect patients with or without concurrent pulmonary fibrosis and/or bone marrow failure [172,173,174]. Occasionally, liver disease may present as severe unresponsive hypoxemia (hepatopulmonary syndrome) [161,175,176]. Taking a careful clinical and family history focused on extrapulmonary manifestations of STS can provide important prognostic information in patients with IPF, as this is associated with shorter survival [177]. Patients who are carriers of TRG mutations should be systematically examined for both pulmonary and extrapulmonary disease severity and progression. It is recommended to stop or limit environmental toxic factor exposure, including tobacco smoke, alcohol, and professional exposures.
Patients who are carriers of TRG mutations have so far been treated following the recommendations for patients with sporadic IPF: “a conditional YES for antifibrotics” [33,178,179,180]. Androgens, including danazol, are under investigation [181,182]. The possibility of telomerase gene therapies, the potential for CRISPR editing to lengthen telomeres, or other approaches to target molecular defects associated with aging and TRG mutations have been examined but are so far restricted by great obstacles and challenges that probably only a deeper understanding of fundamental lung biology could overcome in the future [127,183].

8. Interferonopathies

8.1. STING-Associated Vasculopathy with Onset in Infancy

(SAVI) is an auto-inflammatory monogenic disease related to heterozygous gain-of-function mutations in STING1, an encoding stimulator of interferon genes (STING). Patients present early in life with features of systemic and peripheral vascular inflammation, vascular and tissue damage (nail dystrophy and gangrene of fingers or toes, nasal septum perforation), low-titer autoantibodies, and f-ILD (Table 1). Pathobiologically, the stimulation of STING activates endothelial cells and induces upregulation of interferon-response genes, apoptosis-pathway genes, and endothelial cell death [184,185,186,187].
Aberrant activation of the STING pathway is also detected in COPA syndrome, characterized by dysregulation of the innate and adaptive immune response. Clinically, symptoms occur with varying severity, including ILD with or without pulmonary hemorrhage, inflammatory arthritis, or immune-mediated kidney disease; autoantibodies are evident in most young patients [188]. COPA partially overlaps with SAVI in features such as severe systemic inflammation, recurrent fevers, ILD, early onset in life, and a predominant constitutive type I IFN gene activation. It has been shown that dominant autosomal loss-of-function mutations in the COPA gene, which encodes the α-subunit (COPα, COPA) of the coatomer complex I (COPI) lead in COPA deficiency and cause disease through defective retrograde trafficking, specifically the lost Golgi-to-ER retrieval of immune signaling protein STING pathway (Table 1) [189,190,191,192]. An increased specific “interferon signature” associating IFNa dosage and expression of IFN stimulated genes (ISGs) could be found in blood samples of patients with STING1 or COPA mutations.

8.2. Hermansky–Pudlak Syndrome

Hermansky–Pudlak syndrome (HPS), first described in 1959, is an inherited multisystem autosomal recessive disease characterized by oculocutaneous albinism, bleeding diathesis, inflammatory bowel disease, and pulmonary fibrosis [193,194]. There are 10 genes implicated in HPS: AP3B1, HPS1, HPS3, HPS4, HPS5, HPS6, and less commonly, AP3D1, BLOC1S3, BLOC1S6, and DTNBP1, inherited in an autosomal recessive manner. Pulmonary fibrosis has so far been described in HPS1, HPS4, and AP3B1 genes (Table 1) [195,196,197,198,199]. Hermansky–Pudlak syndrome type 2 manifests with fibrosing lung disease early in childhood [200]. The pathobiology of the syndrome is related to the defective formation of lysosome-related organelles and abnormal intracellular vesicle trafficking due to mutations in genes regulating protein complexes such as BLOC-1, BLOC-2, BLOC-3, and the AP-3 complex, manifesting in various organs and cells including AEC2; endoplasmic reticulum stress, impaired autophagy, defective surfactant secretion, and altered phospholipid content in association with the dysregulation of alveolar macrophages, triggering the development of lung fibrosis [201,202,203,204,205,206,207,208].

9. Clinical Implications of Other Rare Monogenic Forms of Pulmonary Fibrosis

Mutations in genes related to rare syndromic forms of IPF with characteristic extrapulmonary manifestations have historically contributed to the understanding of its heritability and may add to the search for its pathogenesis, unraveling specific pathobiologic pathways [164,209]. The major clinical implications of the genetic analysis in case of extrapulmonary findings consistent with other syndromic forms of pulmonary fibrosis, such as Hermansky–Pudlak, STING, and COPA syndromes, consist mainly in the documentation of the diagnosis and the initiation of specific treatments. The clinical utility of genetic analysis in that case is associated with personalized treatment approaches as well as multidisciplinary collaboration for management by expert centers. The participation in international registries and multicenter clinical trials is the only way that further expert knowledge could be gained for these rare and very challenging diseases [210,211,212,213,214].

10. Common Variants in IPF and MUC5B rs35705950

The development of advanced genetic techniques such as genome wide association studies (GWAS) allows the examination of the association of hundreds of variants and the risk of developing IPF. So far, more than 20 mutations in genes implicated in host defense, cell–cell adhesion, and DNA repair have been shown to genetically predispose to the disease, demonstrating that IPF could be considered not only a monogenic but also a polygenic disease with a lot of yet unidentified variants implicated in the susceptibility of the disease and that genetic susceptibility regards not only familial but also sporadic cases [29,30,215]. The common single nucleotide polymorphism (SNP) in the promoter region of the gene MUC5B that encodes muc5B protein (MUC5B rs 35705950T allele) as well as the rest of the common SNP associated with IPF is shown in Figure 2 [30,55,216,217,218,219,220,221,222].
In 2011, a common SNP in the promoter region of the gene MUC5B that encodes muc5B protein (MUC5B rs 35705950T allele) was reported as the strongest polygenic risk factor for the development of familial PF and IPF [223]. MUC5B rs35705950 T risk allele carriership increases the odds ratio in heterozygous and homozygous patients by 6.8 and 20.8 for familial ILD and 9.0 and 21.8 for IPF, respectively [223]. This association was found by several independent groups of researchers initially in IPF patients and later on in other forms of progressive fibrotic lung disease, such as rheumatoid arthritis-UIP-ILD, fibrotic hypersensitivity pneumonitis, asbestosis, and interstitial lung abnormalities (ILAs) [224,225,226,227,228,229,230,231,232]. In this context, pulmonary fibrosis mechanisms relate to the altered biology of the bronchial secretory cell. MUC5B rs35705950 T risk allele carriership increases the expression of mucin5B in the distal lung, causes mucociliary dysfunction and ER stress, and triggers a vicious cycle of injury/repair and regeneration at the bronchoalveolar junction leading to fibrosis [233,234,235,236,237,238,239,240,241].

11. Clinical Implications of Single Nucleotide Polymorphism and MUC5B rs35705950 T Risk Allele Carriership

Genome-wide association studies are constantly revealing common SNPs associated with an increased risk for IPF that could be considered in a polygenic transmission model of IPF in contrast to the monogenic one so far described. The clinical implications of those SNPs are not extensively studied with the exception of the MUC5B rs35705950 T risk allele and the SNP in TOLLIP [242]. MUC5B rs35705950 T risk allele carriership is related in both sporadic and familial PF with confident UIP patterns both on HRCT and on histology [243,244,245,246]. In asymptomatic relatives of patients with familial PF as well as in any other individual incidentally found to have ILAs, carriership of MUC5B rs35705950 T risk allele is significantly associated not only to the development but also to the progression of PF [230,231,247,248]. In patients with rheumatoid arthritis, MUC5B rs35705950 T risk allele carriership has recently been shown to have a significant contribution to the prediction of subclinical RA-ILD disease [249]. Little evidence exists so far that SNPs in common variants could be used to stratify IPF patients and predict response to treatment [250]. Post-hoc analysis in the PANTHER-IPF trial showed that rs3750920 (TOLLIP) TT genotype affected positively the response to N-acetylcysteine (NAC) providing evidence that in a precision-medicine era, genotype-stratified prospective clinical trials should be conducted before any recommendation for therapeutic options in IPF [251]. Despite reports identifying common SNPs and rare missense variants associated with ILD, there are no large multicenter studies that assess the clinical value of screening for rare ILD-linked genetic variants in patients with sporadic IPF.

12. Special Considerations

12.1. Children Diagnosed with ILD Reaching Adult Life

Special consideration is due to genetics in children’s interstitial lung disease (chILD). The range of diseases in that case becomes even broader, including, in addition to SRG- and TRG-related diseases, many rare and ultra-rare entities, such as surfactant protein catabolic disorders related to CSFRA, CSFRB, and OAS1 mutations; syndromic forms of ILD due to intergrin mutations; and chILD related to filamin A variants [25,252,253,254,255]. The scope of the present review could not embrace such a broad spectrum and focuses only on adult disease. However, as the quality of care ameliorates for chILD, survival into adulthood allows a number of those patients to transfer from pediatric to adult care and engages adult pulmonologists to meet their needs [256]. A recent survey demonstrated that the availability of data regarding the characteristics of these patients as well as their transitional care are scarce and tailored the way for immediate action through a dedicated ERS task force whose results are highly anticipated in the near future [257,258].

12.2. Genetic Testing for f-ILDs in Everyday Clinical Practice

At an international level, genetic testing for f-ILDS is not clinically available in every country, nor could it be available in each institution given the increased costs and specialized knowledge that are required. Even more importantly, the interpretation of the results of genetic analysis, especially the pathogenic role of every variant that is identified is a very demanding, time-consuming process that only expert centers can perform. Even in that case, multidisciplinary discussion and international collaboration are often needed. Genetic testing is complete only when genetic counselling is offered to the patients and their families in order to assimilate the clinical, psychological, social, and familial implications of genetic predisposition to disease and adapt accordingly. Based on expert opinion, the clinical indications for which genetic testing may be considered with a high probability of identifying a pathogenic mutation include patients with familial PF, patients with personal or family features of telomeropathy, syndromic forms of the disease, early-onset pulmonary fibrosis without any other obvious explanation, and asymptomatic relatives of patients with known pathogenic disease-causing variants. Genetic testing could also be considered in families where pulmonary fibrosis and adenocarcinoma of the lung segregate in many family members across generations and in patients with ILAs, as well as before lung transplantation. Genetic testing for sporadic PF is not currently recommended [64,65,259].
Official guidelines are still missing; however, expert opinion regarding the examination of relatives of IPF-associated pathogenic variants suggests an initial evaluation at a specialized multidisciplinary outpatient clinic and systematic evaluation thereafter, depending on the individual findings in each case. Genetic analysis may be performed only at the request of a relative who is examined after written informed consent. Depending on the country in question’s legislation, only adult asymptomatic relatives may benefit from a genetic testing. An action plan for the avoidance of toxic exposures is proposed and special care is undertaken depending on the presence or absence of a specific disease. If the genetic analysis confirms the presence of a mutation in an asymptomatic and healthy relative, a regular follow-up is proposed every 3–5 years, unless symptoms suggestive of progressive disease develop earlier [64]. The above recommendations are based on the findings of elaborate cohort studies including first-degree asymptomatic relatives of patients with familial f-ILD that have demonstrated the presence of substantial (25%) subclinical disease [63,260,261,262,263,264]. Early diagnosis of any progressive f-ILD in the familial context could lead to earlier treatment, but whether early treatment translates to improved outcomes is currently unknown and needs to be further explored [265].

13. Conclusions—Future Perspective

In the future, the integration of genetic evaluation in the diagnostic algorithm of IPF next to long-standing traditional modalities such as HRCT, BAL, and lung biopsy could support precision medicine and guide the multidisciplinary discussion team to personalized diagnostic and therapeutic approaches. The detection and referral of at-risk populations, including asymptomatic relatives of carriers of known mutations for early recognition of the disease, could optimize clinical outcomes in a currently lately documented and incurable disease. Identifying interstitial lung disease with a genetic background in pediatric populations that survive into adult life could have a tremendous impact on timely diagnosis ensuring a better quality of life, targeted therapies and a successful transition to expert adult care. Genetic counseling regarding the mode of inheritance and the risk to offspring and the other family members provided to all IPF patients with a genetic risk should help informed decision making. Pre-natal and pre-implantation genetic testing bear significant ethical implications and genotype identification and family history are often insufficient to predict the course of disease in an individual and therefore should be handled with caution.
In the future, the decoding of missing heritability in IPF could further advance our understanding of the disease and offer the opportunity to delay its progression and increase survival. The collaborative work and solidarity of scientific experts and clinicians all over the world regarding the role of genetics in IPF patients is the way to safeguard the tremendous knowledge that genetic science offers, transforming it into healing wisdom.

Author Contributions

Conceptualization, S.A.P. and E.D.M.; methodology, S.A.P., E.D.M., C.K., I.B. and A.B.; software, S.A.P., M.K. and E.D.M.; validation, C.K., R.B., N.N., M.G., P.D. and B.C.; formal analysis, S.A.P., L.K., V.A. and E.D.M.; investigation, S.A.P., C.K., R.B., M.K., V.A. and E.D.M.; resources, L.K.; data curation, S.A.P., C.K. and E.D.M.; writing—original draft preparation, S.A.P. and E.D.M.; writing—review and editing, C.K., R.B., I.B., N.N., A.B., M.G., P.D. and B.C.; visualization, S.A.P., V.A. and E.D.M.; supervision, E.D.M.; project administration, S.A.P. and E.D.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Raghu, G.; Remy-Jardin, M.; Richeldi, L.; Thomson, C.C.; Inoue, Y.; Johkoh, T.; Kreuter, M.; Lynch, D.A.; Maher, T.M.; Martinez, F.J.; et al. Idiopathic Pulmonary Fibrosis (an Update) and Progressive Pulmonary Fibrosis in Adults: An Official ATS/ERS/JRS/ALAT Clinical Practice Guideline. Am. J. Respir. Crit. Care Med. 2022, 205, e18–e47. [Google Scholar] [CrossRef] [PubMed]
  2. Valenzuela, C.; Cottin, V. Epidemiology and real-life experience in progressive pulmonary fibrosis. Curr. Opin. Pulm. Med. 2022, 28, 407–413. [Google Scholar] [CrossRef] [PubMed]
  3. Wijsenbeek, M.; Cottin, V. Spectrum of Fibrotic Lung Diseases. N. Engl. J. Med. 2020, 383, 958–968. [Google Scholar] [CrossRef] [PubMed]
  4. Wijsenbeek, M.; Suzuki, A.; Maher, T.M. Interstitial lung diseases. Lancet 2022, 400, 769–786. [Google Scholar] [CrossRef] [PubMed]
  5. Cottin, V.; Hirani, N.A.; Hotchkin, D.L.; Nambiar, A.M.; Ogura, T.; Otaola, M.; Skowasch, D.; Park, J.S.; Poonyagariyagorn, H.K.; Wuyts, W.; et al. Presentation, diagnosis and clinical course of the spectrum of progressive-fibrosing interstitial lung diseases. Eur. Respir. Rev. 2018, 27, 180076. [Google Scholar] [CrossRef] [Green Version]
  6. Pugashetti, J.V.; Adegunsoye, A.; Wu, Z.; Lee, C.T.; Srikrishnan, A.; Ghodrati, S.; Vo, V.; Renzoni, E.A.; Wells, A.U.; Garcia, C.K.; et al. Validation of Proposed Criteria for Progressive Pulmonary Fibrosis. Am. J. Respir. Crit. Care Med. 2022. [Google Scholar] [CrossRef]
  7. Galioto, F.; Palmucci, S.; Astuti, G.M.; Vancheri, A.; Distefano, G.; Tiralongo, F.; Libra, A.; Cusumano, G.; Basile, A.; Vancheri, C. Complications in Idiopathic Pulmonary Fibrosis: Focus on Their Clinical and Radiological Features. Diagnostics 2020, 10, 450. [Google Scholar] [CrossRef]
  8. Baratella, E.; Ruaro, B.; Giudici, F.; Wade, B.; Santagiuliana, M.; Salton, F.; Confalonieri, P.; Simbolo, M.; Scarpa, A.; Tollot, S.; et al. Evaluation of Correlations between Genetic Variants and High-Resolution Computed Tomography Patterns in Idiopathic Pulmonary Fibrosis. Diagnostics 2021, 11, 762. [Google Scholar] [CrossRef]
  9. Kropski, J.A.; Lawson, W.E.; Young, L.R.; Blackwell, T.S. Genetic studies provide clues on the pathogenesis of idiopathic pulmonary fibrosis. Dis. Model Mech. 2013, 6, 9–17. [Google Scholar] [CrossRef] [Green Version]
  10. Borie, R.; Le Guen, P.; Ghanem, M.; Taillé, C.; Dupin, C.; Dieudé, P.; Kannengiesser, C.; Crestani, B. The genetics of interstitial lung diseases. Eur. Respir. Rev. 2019, 28, 190053. [Google Scholar] [CrossRef]
  11. Zhang, D.; Newton, C.A. Familial Pulmonary Fibrosis: Genetic Features and Clinical Implications. Chest 2021, 160, 1764–1773. [Google Scholar] [CrossRef] [PubMed]
  12. Schimmelpennink, M.C.; Meek, D.B.; Vorselaars, A.D.M.; Langezaal, L.C.M.; van Moorsel, C.H.M.; van der Vis, J.J.; Veltkamp, M.; Grutters, J.C. Characterization of the PF-ILD phenotype in patients with advanced pulmonary sarcoidosis. Respir. Res. 2022, 23, 169. [Google Scholar] [CrossRef] [PubMed]
  13. O’Callaghan, M.; Bonella, F.; McCarthy, C. Unclassifiable, or simply unclassified interstitial lung disease? Curr. Opin. Pulm. Med. 2021, 27, 405–413. [Google Scholar] [CrossRef]
  14. Poellinger, A.; Berezowska, S.; Myers, J.L.; Huber, A.; Funke-Chambour, M.; Guler, S.; Geiser, T.; Harari, S.; Caminati, A.; Zompatori, M.; et al. The Octopus Sign-A New HRCT Sign in Pulmonary Langerhans Cell Histiocytosis. Diagnostics 2022, 12, 937. [Google Scholar] [CrossRef] [PubMed]
  15. Katzenstein, A.L.; Mukhopadhyay, S.; Myers, J.L. Diagnosis of usual interstitial pneumonia and distinction from other fibrosing interstitial lung diseases. Hum. Pathol. 2008, 39, 1275–1294. [Google Scholar] [CrossRef] [PubMed]
  16. Lynch, D.A.; Sverzellati, N.; Travis, W.D.; Brown, K.K.; Colby, T.V.; Galvin, J.R.; Goldin, J.G.; Hansell, D.M.; Inoue, Y.; Johkoh, T.; et al. Diagnostic criteria for idiopathic pulmonary fibrosis: A Fleischner Society White Paper. Lancet Respir. Med. 2018, 6, 138–153. [Google Scholar] [CrossRef] [PubMed]
  17. Adegunsoye, A.; Vij, R.; Noth, I. Integrating Genomics Into Management of Fibrotic Interstitial Lung Disease. Chest 2019, 155, 1026–1040. [Google Scholar] [CrossRef]
  18. Newton, C.A.; Batra, K.; Torrealba, J.; Kozlitina, J.; Glazer, C.S.; Aravena, C.; Meyer, K.; Raghu, G.; Collard, H.R.; Garcia, C.K. Telomere-related lung fibrosis is diagnostically heterogeneous but uniformly progressive. Eur. Respir. J. 2016, 48, 1710–1720. [Google Scholar] [CrossRef] [Green Version]
  19. Borie, R.; Tabèze, L.; Thabut, G.; Nunes, H.; Cottin, V.; Marchand-Adam, S.; Prevot, G.; Tazi, A.; Cadranel, J.; Mal, H.; et al. Prevalence and characteristics of TERT and TERC mutations in suspected genetic pulmonary fibrosis. Eur. Respir. J. 2016, 48, 1721–1731. [Google Scholar] [CrossRef] [Green Version]
  20. Loyd, J.E. Pulmonary fibrosis in families. Am. J. Respir. Cell Mol. Biol. 2003, 29, S47–S50. [Google Scholar]
  21. Steele, M.P.; Speer, M.C.; Loyd, J.E.; Brown, K.K.; Herron, A.; Slifer, S.H.; Burch, L.H.; Wahidi, M.M.; Phillips, J.A., 3rd; Sporn, T.A.; et al. Clinical and pathologic features of familial interstitial pneumonia. Am. J. Respir. Crit. Care Med. 2005, 172, 1146–1152. [Google Scholar] [CrossRef] [PubMed]
  22. Calado, R.T.; Young, N.S. Telomere diseases. N. Engl. J. Med. 2009, 361, 2353–2365. [Google Scholar] [CrossRef] [PubMed]
  23. Tummala, H.; Walne, A.; Dokal, I. The biology and management of dyskeratosis congenita and related disorders of telomeres. Expert Rev. Hematol. 2022, 15, 685–696. [Google Scholar] [CrossRef] [PubMed]
  24. Borie, R.; Kannengiesser, C.; Dupin, C.; Debray, M.P.; Cazes, A.; Crestani, B. Impact of genetic factors on fibrosing interstitial lung diseases. Incidence and clinical presentation in adults. Presse Med. 2020, 49, 104024. [Google Scholar] [CrossRef]
  25. Laenger, F.P.; Schwerk, N.; Dingemann, J.; Welte, T.; Auber, B.; Verleden, S.; Ackermann, M.; Mentzer, S.J.; Griese, M.; Jonigk, D. Interstitial lung disease in infancy and early childhood: A clinicopathological primer. Eur. Respir. Rev. 2022, 31, 210251. [Google Scholar] [CrossRef]
  26. Griese, M. Chronic interstitial lung disease in children. Eur. Respir. Rev. 2018, 27, 170100. [Google Scholar] [CrossRef] [Green Version]
  27. Nogee, L.M.; Hamvas, A. The past and future of genetics in pulmonary disease: You can teach an old dog new tricks. Pediatr. Pulmonol. 2020, 55, 1789–1793. [Google Scholar] [CrossRef]
  28. Kaur, A.; Mathai, S.K.; Schwartz, D.A. Genetics in Idiopathic Pulmonary Fibrosis Pathogenesis, Prognosis, and Treatment. Front. Med. 2017, 4, 154. [Google Scholar] [CrossRef] [Green Version]
  29. Allen, R.J.; Guillen-Guio, B.; Oldham, J.M.; Ma, S.F.; Dressen, A.; Paynton, M.L.; Kraven, L.M.; Obeidat, M.; Li, X.; Ng, M.; et al. Genome-Wide Association Study of Susceptibility to Idiopathic Pulmonary Fibrosis. Am. J. Respir. Crit. Care Med. 2020, 201, 564–574. [Google Scholar] [CrossRef]
  30. Allen, R.J.; Stockwell, A.; Oldham, J.M.; Guillen-Guio, B.; Schwartz, D.A.; Maher, T.M.; Flores, C.; Noth, I.; Yaspan, B.L.; Jenkins, R.G.; et al. Genome-wide association study across five cohorts identifies five novel loci associated with idiopathic pulmonary fibrosis. Thorax 2022, 77, 829–833. [Google Scholar] [CrossRef]
  31. Moore, C.; Blumhagen, R.Z.; Yang, I.V.; Walts, A.; Powers, J.; Walker, T.; Bishop, M.; Russell, P.; Vestal, B.; Cardwell, J.; et al. Resequencing Study Confirms That Host Defense and Cell Senescence Gene Variants Contribute to the Risk of Idiopathic Pulmonary Fibrosis. Am. J. Respir. Crit. Care Med. 2019, 200, 199–208. [Google Scholar] [CrossRef] [PubMed]
  32. Manali, E.D.; Kannengiesser, C.; Borie, R.; Ba, I.; Bouros, D.; Markopoulou, A.; Antoniou, K.; Kolilekas, L.; Papaioannou, A.I.; Tzilas, V.; et al. Genotype-Phenotype Relationships in Inheritable Idiopathic Pulmonary Fibrosis: A Greek National Cohort Study. Respiration 2022, 101, 531–543. [Google Scholar] [CrossRef] [PubMed]
  33. Dressen, A.; Abbas, A.R.; Cabanski, C.; Reeder, J.; Ramalingam, T.R.; Neighbors, M.; Bhangale, T.R.; Brauer, M.J.; Hunkapiller, J.; Reeder, J.; et al. Analysis of protein-altering variants in telomerase genes and their association with MUC5B common variant status in patients with idiopathic pulmonary fibrosis: A candidate gene sequencing study. Lancet Respir. Med. 2018, 6, 603–614. [Google Scholar] [CrossRef] [PubMed]
  34. Zhang, D.; Newton, C.A.; Wang, B.; Povysil, G.; Noth, I.; Martinez, F.J.; Raghu, G.; Goldstein, D.; Kim Garcia, C. Utility of whole genome sequencing in assessing risk and clinically-relevant outcomes for pulmonary fibrosis. Eur. Respir. J. 2022, 2200577. [Google Scholar] [CrossRef] [PubMed]
  35. Kraven, L.M.; Taylor, A.R.; Molyneaux, P.L.; Maher, T.M.; McDonough, J.E.; Mura, M.; Yang, I.V.; Schwartz, D.A.; Huang, Y.; Noth, I.; et al. Cluster analysis of transcriptomic datasets to identify endotypes of idiopathic pulmonary fibrosis. Thorax 2022. [Google Scholar] [CrossRef]
  36. Juan-Guardela, B.M.; Herazo-Maya, J.D. Immunity, Ciliated Epithelium, and Mortality: Are We Ready to Identify Idiopathic Pulmonary Fibrosis Endotypes With Prognostic Significance? Chest 2022, 161, 1440–1441. [Google Scholar] [CrossRef]
  37. Moss, B.J.; Ryter, S.W.; Rosas, I.O. Pathogenic Mechanisms Underlying Idiopathic Pulmonary Fibrosis. Annu. Rev. Pathol. 2022, 17, 515–546. [Google Scholar] [CrossRef]
  38. Martinez, F.J.; Collard, H.R.; Pardo, A.; Raghu, G.; Richeldi, L.; Selman, M.; Swigris, J.J.; Taniguchi, H.; Wells, A.U. Idiopathic pulmonary fibrosis. Nat. Rev. Dis. Primers 2017, 3, 17074. [Google Scholar] [CrossRef]
  39. Distler, J.H.W.; Györfi, A.H.; Ramanujam, M.; Whitfield, M.L.; Königshoff, M.; Lafyatis, R. Shared and distinct mechanisms of fibrosis. Nat. Rev. Rheumatol. 2019, 15, 705–730. [Google Scholar] [CrossRef]
  40. Lagares, D.; Hinz, B. Animal and Human Models of Tissue Repair and Fibrosis: An Introduction. Methods Mol. Biol. 2021, 2299, 277–290. [Google Scholar] [CrossRef]
  41. Vannella, K.M.; Wynn, T.A. Mechanisms of Organ Injury and Repair by Macrophages. Annu. Rev. Physiol. 2017, 79, 593–617. [Google Scholar] [CrossRef] [PubMed]
  42. Wynn, T.A.; Ramalingam, T.R. Mechanisms of fibrosis: Therapeutic translation for fibrotic disease. Nat. Med. 2012, 18, 1028–1040. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Eming, S.A.; Martin, P.; Tomic-Canic, M. Wound repair and regeneration: Mechanisms, signaling, and translation. Sci. Transl. Med. 2014, 6, 265sr266. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Kobayashi, Y.; Tata, A.; Konkimalla, A.; Katsura, H.; Lee, R.F.; Ou, J.; Banovich, N.E.; Kropski, J.A.; Tata, P.R. Persistence of a regeneration-associated, transitional alveolar epithelial cell state in pulmonary fibrosis. Nat. Cell Biol. 2020, 22, 934–946. [Google Scholar] [CrossRef] [PubMed]
  45. Leslie, K.O. Idiopathic pulmonary fibrosis may be a disease of recurrent, tractional injury to the periphery of the aging lung: A unifying hypothesis regarding etiology and pathogenesis. Arch. Pathol Lab. Med. 2012, 136, 591–600. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Wu, H.; Yu, Y.; Huang, H.; Hu, Y.; Fu, S.; Wang, Z.; Shi, M.; Zhao, X.; Yuan, J.; Li, J.; et al. Progressive pulmonary fibrosis is caused by elevated mechanical tension on alveolar stem cells. Cell 2021, 184, 845–846. [Google Scholar] [CrossRef] [PubMed]
  47. Froese, A.R.; Shimbori, C.; Bellaye, P.S.; Inman, M.; Obex, S.; Fatima, S.; Jenkins, G.; Gauldie, J.; Ask, K.; Kolb, M. Stretch-induced Activation of Transforming Growth Factor-β1 in Pulmonary Fibrosis. Am. J. Respir. Crit. Care Med. 2016, 194, 84–96. [Google Scholar] [CrossRef]
  48. Smith, M.L. The histologic diagnosis of usual interstitial pneumonia of idiopathic pulmonary fibrosis. Where we are and where we need to go. Mod. Pathol. 2022, 35, 8–14. [Google Scholar] [CrossRef]
  49. Tomos, I.P.; Tzouvelekis, A.; Aidinis, V.; Manali, E.D.; Bouros, E.; Bouros, D.; Papiris, S.A. Extracellular matrix remodeling in idiopathic pulmonary fibrosis. It is the ‘bed’ that counts and not ‘the sleepers’. Expert Rev. Respir. Med. 2017, 11, 299–309. [Google Scholar] [CrossRef]
  50. Selman, M.; Pardo, A. From pulmonary fibrosis to progressive pulmonary fibrosis: A lethal pathobiological jump. Am. J. Physiol. Lung Cell Mol. Physiol. 2021, 321, L600–L607. [Google Scholar] [CrossRef]
  51. Selman, M.; Pardo, A. When things go wrong: Exploring possible mechanisms driving the progressive fibrosis phenotype in interstitial lung diseases. Eur. Respir. J. 2021, 58, 2004507. [Google Scholar] [CrossRef] [PubMed]
  52. Thannickal, V.J.; Murthy, M.; Balch, W.E.; Chandel, N.S.; Meiners, S.; Eickelberg, O.; Selman, M.; Pardo, A.; White, E.S.; Levy, B.D.; et al. Blue journal conference. Aging and susceptibility to lung disease. Am. J. Respir. Crit. Care Med. 2015, 191, 261–269. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Moimas, S.; Salton, F.; Kosmider, B.; Ring, N.; Volpe, M.C.; Bahmed, K.; Braga, L.; Rehman, M.; Vodret, S.; Graziani, M.L.; et al. miR-200 family members reduce senescence and restore idiopathic pulmonary fibrosis type II alveolar epithelial cell transdifferentiation. ERJ Open Res. 2019, 5, 869–880. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Leavy, O.C.; Ma, S.F.; Molyneaux, P.L.; Maher, T.M.; Oldham, J.M.; Flores, C.; Noth, I.; Jenkins, R.G.; Dudbridge, F.; Wain, L.V.; et al. Proportion of Idiopathic Pulmonary Fibrosis Risk Explained by Known Common Genetic Loci in European Populations. Am. J. Respir. Crit. Care Med. 2021, 203, 775–778. [Google Scholar] [CrossRef]
  55. Allen, R.J.; Porte, J.; Braybrooke, R.; Flores, C.; Fingerlin, T.E.; Oldham, J.M.; Guillen-Guio, B.; Ma, S.F.; Okamoto, T.; John, A.E.; et al. Genetic variants associated with susceptibility to idiopathic pulmonary fibrosis in people of European ancestry: A genome-wide association study. Lancet Respir. Med. 2017, 5, 869–880. [Google Scholar] [CrossRef] [Green Version]
  56. Witte, J.S.; Visscher, P.M.; Wray, N.R. The contribution of genetic variants to disease depends on the ruler. Nat. Rev. Genet. 2014, 15, 765–776. [Google Scholar] [CrossRef] [Green Version]
  57. Copeland, C.R.; Donnelly, E.F.; Mehrad, M.; Ding, G.; Markin, C.R.; Douglas, K.; Wu, P.; Cogan, J.D.; Young, L.R.; Bartholmai, B.J.; et al. The Association between Exposures and Disease Characteristics in Familial Pulmonary Fibrosis. Ann. Am. Thorac. Soc. 2022. [Google Scholar] [CrossRef]
  58. Hunninghake, G.M.; Goldin, J.G.; Kadoch, M.A.; Kropski, J.A.; Rosas, I.O.; Wells, A.U.; Yadav, R.; Lazarus, H.M.; Abtin, F.G.; Corte, T.J.; et al. Detection and Early Referral of Patients With Interstitial Lung Abnormalities: An Expert Survey Initiative. Chest 2022, 161, 470–482. [Google Scholar] [CrossRef]
  59. Hunninghake, G.M.; Quesada-Arias, L.D.; Carmichael, N.E.; Martinez Manzano, J.M.; Poli De Frías, S.; Baumgartner, M.A.; DiGianni, L.; Gampala-Sagar, S.N.; Leone, D.A.; Gulati, S.; et al. Interstitial Lung Disease in Relatives of Patients with Pulmonary Fibrosis. Am. J. Respir. Crit. Care Med. 2020, 201, 1240–1248. [Google Scholar] [CrossRef]
  60. Putman, R.K.; Rosas, I.O.; Hunninghake, G.M. Genetics and early detection in idiopathic pulmonary fibrosis. Am. J. Respir. Crit. Care Med. 2014, 189, 770–778. [Google Scholar] [CrossRef] [Green Version]
  61. Salisbury, M.L.; Hewlett, J.C.; Ding, G.; Markin, C.R.; Douglas, K.; Mason, W.; Guttentag, A.; Phillips, J.A., 3rd; Cogan, J.D.; Reiss, S.; et al. Development and Progression of Radiologic Abnormalities in Individuals at Risk for Familial Interstitial Lung Disease. Am. J. Respir. Crit. Care Med. 2020, 201, 1230–1239. [Google Scholar] [CrossRef] [PubMed]
  62. Kropski, J.A.; Young, L.R.; Cogan, J.D.; Mitchell, D.B.; Lancaster, L.H.; Worrell, J.A.; Markin, C.; Liu, N.; Mason, W.R.; Fingerlin, T.E.; et al. Genetic Evaluation and Testing of Patients and Families with Idiopathic Pulmonary Fibrosis. Am. J. Respir. Crit. Care Med. 2017, 195, 1423–1428. [Google Scholar] [CrossRef] [PubMed]
  63. Kropski, J.A.; Pritchett, J.M.; Zoz, D.F.; Crossno, P.F.; Markin, C.; Garnett, E.T.; Degryse, A.L.; Mitchell, D.B.; Polosukhin, V.V.; Rickman, O.B.; et al. Extensive phenotyping of individuals at risk for familial interstitial pneumonia reveals clues to the pathogenesis of interstitial lung disease. Am. J. Respir. Crit. Care Med. 2015, 191, 417–426. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Borie, R.; Kannengiesser, C.; Sicre de Fontbrune, F.; Gouya, L.; Nathan, N.; Crestani, B. Management of suspected monogenic lung fibrosis in a specialised centre. Eur. Respir. Rev. 2017, 26, 160122. [Google Scholar] [CrossRef]
  65. Borie, R.; Kannengiesser, C.; Gouya, L.; Dupin, C.; Amselem, S.; Ba, I.; Bunel, V.; Bonniaud, P.; Bouvry, D.; Cazes, A.; et al. Pilot experience of multidisciplinary team discussion dedicated to inherited pulmonary fibrosis. Orphanet. J. Rare Dis. 2019, 14, 280. [Google Scholar] [CrossRef] [Green Version]
  66. Eckardt, N.A.; Birchler, J.A.; Meyers, B.C. Focus on plant genetics: Celebrating Gregor Mendel’s 200th birth anniversary. Plant Cell 2022, 34, 2453–2454. [Google Scholar] [CrossRef]
  67. Davies, K. After the genome: DNA and human disease. Cell 2001, 104, 465–467. [Google Scholar] [CrossRef] [Green Version]
  68. Jasny, B.R.; Kennedy, D. The human genome. Science 2001, 291, 1153. [Google Scholar] [CrossRef]
  69. Peabody, J.W.; Peabody, J.W., Jr.; Hayes, E.W.; Hayes, E.W., Jr. Idiopathic pulmonary fibrosis; its occurrence in identical twin sisters. Dis. Chest 1950, 18, 330–344. [Google Scholar] [CrossRef]
  70. Marshall, R.P.; McAnulty, R.J.; Laurent, G.J. The pathogenesis of pulmonary fibrosis: Is there a fibrosis gene? Int. J. Biochem. Cell Biol. 1997, 29, 107–120. [Google Scholar] [CrossRef]
  71. Nogee, L.M.; Dunbar, A.E., 3rd; Wert, S.E.; Askin, F.; Hamvas, A.; Whitsett, J.A. A mutation in the surfactant protein C gene associated with familial interstitial lung disease. N. Engl. J. Med. 2001, 344, 573–579. [Google Scholar] [CrossRef] [PubMed]
  72. Thomas, A.Q.; Lane, K.; Phillips, J., 3rd; Prince, M.; Markin, C.; Speer, M.; Schwartz, D.A.; Gaddipati, R.; Marney, A.; Johnson, J.; et al. Heterozygosity for a surfactant protein C gene mutation associated with usual interstitial pneumonitis and cellular nonspecific interstitial pneumonitis in one kindred. Am. J. Respir. Crit. Care Med. 2002, 165, 1322–1328. [Google Scholar] [CrossRef] [PubMed]
  73. Ono, S.; Tanaka, T.; Ishida, M.; Kinoshita, A.; Fukuoka, J.; Takaki, M.; Sakamoto, N.; Ishimatsu, Y.; Kohno, S.; Hayashi, T.; et al. Surfactant protein C G100S mutation causes familial pulmonary fibrosis in Japanese kindred. Eur. Respir. J. 2011, 38, 861–869. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. van Moorsel, C.H.; van Oosterhout, M.F.; Barlo, N.P.; de Jong, P.A.; van der Vis, J.J.; Ruven, H.J.; van Es, H.W.; van den Bosch, J.M.; Grutters, J.C. Surfactant protein C mutations are the basis of a significant portion of adult familial pulmonary fibrosis in a dutch cohort. Am. J. Respir. Crit. Care Med. 2010, 182, 1419–1425. [Google Scholar] [CrossRef] [PubMed]
  75. Wang, Y.; Kuan, P.J.; Xing, C.; Cronkhite, J.T.; Torres, F.; Rosenblatt, R.L.; DiMaio, J.M.; Kinch, L.N.; Grishin, N.V.; Garcia, C.K. Genetic defects in surfactant protein A2 are associated with pulmonary fibrosis and lung cancer. Am. J. Hum. Genet. 2009, 84, 52–59. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Nathan, N.; Giraud, V.; Picard, C.; Nunes, H.; Dastot-Le Moal, F.; Copin, B.; Galeron, L.; De Ligniville, A.; Kuziner, N.; Reynaud-Gaubert, M.; et al. Germline SFTPA1 mutation in familial idiopathic interstitial pneumonia and lung cancer. Hum. Mol Genet. 2016, 25, 1457–1467. [Google Scholar] [CrossRef] [Green Version]
  77. Shulenin, S.; Nogee, L.M.; Annilo, T.; Wert, S.E.; Whitsett, J.A.; Dean, M. ABCA3 gene mutations in newborns with fatal surfactant deficiency. N. Engl. J. Med. 2004, 350, 1296–1303. [Google Scholar] [CrossRef] [Green Version]
  78. Bullard, J.E.; Wert, S.E.; Whitsett, J.A.; Dean, M.; Nogee, L.M. ABCA3 mutations associated with pediatric interstitial lung disease. Am. J. Respir. Crit. Care Med. 2005, 172, 1026–1031. [Google Scholar] [CrossRef]
  79. Kröner, C.; Wittmann, T.; Reu, S.; Teusch, V.; Klemme, M.; Rauch, D.; Hengst, M.; Kappler, M.; Cobanoglu, N.; Sismanlar, T.; et al. Lung disease caused by ABCA3 mutations. Thorax 2017, 72, 213–220. [Google Scholar] [CrossRef] [Green Version]
  80. Manali, E.D.; Legendre, M.; Nathan, N.; Kannengiesser, C.; Coulomb-L’Hermine, A.; Tsiligiannis, T.; Tomos, P.; Griese, M.; Borie, R.; Clement, A.; et al. Bi-allelic missense ABCA3 mutations in a patient with childhood ILD who reached adulthood. ERJ Open Res. 2019, 5, 00066–2019. [Google Scholar] [CrossRef] [Green Version]
  81. Galambos, C.; Levy, H.; Cannon, C.L.; Vargas, S.O.; Reid, L.M.; Cleveland, R.; Lindeman, R.; deMello, D.E.; Wert, S.E.; Whitsett, J.A.; et al. Pulmonary pathology in thyroid transcription factor-1 deficiency syndrome. Am. J. Respir. Crit. Care Med. 2010, 182, 549–554. [Google Scholar] [CrossRef] [PubMed]
  82. Hamvas, A.; Deterding, R.R.; Wert, S.E.; White, F.V.; Dishop, M.K.; Alfano, D.N.; Halbower, A.C.; Planer, B.; Stephan, M.J.; Uchida, D.A.; et al. Heterogeneous pulmonary phenotypes associated with mutations in the thyroid transcription factor gene NKX2-1. Chest 2013, 144, 794–804. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Borie, R.; Funalot, B.; Epaud, R.; Delestrain, C.; Cazes, A.; Gounant, V.; Frija, J.; Debray, M.P.; Zalcman, G.; Crestani, B. NKX2.1 (TTF1) germline mutation associated with pulmonary fibrosis and lung cancer. ERJ Open Res. 2021, 7, 00356-2021. [Google Scholar] [CrossRef] [PubMed]
  84. Nattes, E.; Lejeune, S.; Carsin, A.; Borie, R.; Gibertini, I.; Balinotti, J.; Nathan, N.; Marchand-Adam, S.; Thumerelle, C.; Fauroux, B.; et al. Heterogeneity of lung disease associated with NK2 homeobox 1 mutations. Respir. Med. 2017, 129, 16–23. [Google Scholar] [CrossRef] [PubMed]
  85. van Moorsel, C.H.M.; van der Vis, J.J.; Grutters, J.C. Genetic disorders of the surfactant system: Focus on adult disease. Eur. Respir. Rev. 2021, 30, 200085. [Google Scholar] [CrossRef] [PubMed]
  86. Nogee, L.M. Genetic causes of surfactant protein abnormalities. Curr. Opin. Pediatr. 2019, 31, 330–339. [Google Scholar] [CrossRef]
  87. Lawson, W.E.; Crossno, P.F.; Polosukhin, V.V.; Roldan, J.; Cheng, D.S.; Lane, K.B.; Blackwell, T.R.; Xu, C.; Markin, C.; Ware, L.B.; et al. Endoplasmic reticulum stress in alveolar epithelial cells is prominent in IPF: Association with altered surfactant protein processing and herpesvirus infection. Am. J. Physiol. Lung Cell Mol. Physiol. 2008, 294, L1119–L1126. [Google Scholar] [CrossRef] [Green Version]
  88. Hawkins, A.; Guttentag, S.H.; Deterding, R.; Funkhouser, W.K.; Goralski, J.L.; Chatterjee, S.; Mulugeta, S.; Beers, M.F. A non-BRICHOS SFTPC mutant (SP-CI73T) linked to interstitial lung disease promotes a late block in macroautophagy disrupting cellular proteostasis and mitophagy. Am. J. Physiol. Lung Cell Mol. Physiol. 2015, 308, L33–L47. [Google Scholar] [CrossRef] [Green Version]
  89. Beers, M.F.; Hawkins, A.; Maguire, J.A.; Kotorashvili, A.; Zhao, M.; Newitt, J.L.; Ding, W.; Russo, S.; Guttentag, S.; Gonzales, L.; et al. A nonaggregating surfactant protein C mutant is misdirected to early endosomes and disrupts phospholipid recycling. Traffic 2011, 12, 1196–1210. [Google Scholar] [CrossRef] [Green Version]
  90. Katzen, J.; Wagner, B.D.; Venosa, A.; Kopp, M.; Tomer, Y.; Russo, S.J.; Headen, A.C.; Basil, M.C.; Stark, J.M.; Mulugeta, S.; et al. An SFTPC BRICHOS mutant links epithelial ER stress and spontaneous lung fibrosis. JCI Insight 2019, 4, e126125. [Google Scholar] [CrossRef] [Green Version]
  91. Beers, M.F.; Mulugeta, S. The biology of the ABCA3 lipid transporter in lung health and disease. Cell Tissue Res. 2017, 367, 481–493. [Google Scholar] [CrossRef] [PubMed]
  92. Maitra, M.; Wang, Y.; Gerard, R.D.; Mendelson, C.R.; Garcia, C.K. Surfactant protein A2 mutations associated with pulmonary fibrosis lead to protein instability and endoplasmic reticulum stress. J. Biol. Chem. 2010, 285, 22103–22113. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Takezaki, A.; Tsukumo, S.I.; Setoguchi, Y.; Ledford, J.G.; Goto, H.; Hosomichi, K.; Uehara, H.; Nishioka, Y.; Yasutomo, K. A homozygous SFTPA1 mutation drives necroptosis of type II alveolar epithelial cells in patients with idiopathic pulmonary fibrosis. J. Exp. Med. 2019, 216, 2724–2735. [Google Scholar] [CrossRef]
  94. Carré, A.; Szinnai, G.; Castanet, M.; Sura-Trueba, S.; Tron, E.; Broutin-L’Hermite, I.; Barat, P.; Goizet, C.; Lacombe, D.; Moutard, M.L.; et al. Five new TTF1/NKX2.1 mutations in brain-lung-thyroid syndrome: Rescue by PAX8 synergism in one case. Hum. Mol. Genet. 2009, 18, 2266–2276. [Google Scholar] [CrossRef] [Green Version]
  95. Kim, K.; Shin, D.; Lee, G.; Bae, H. Loss of SP-A in the Lung Exacerbates Pulmonary Fibrosis. Int. J. Mol. Sci. 2022, 23, 5292. [Google Scholar] [CrossRef] [PubMed]
  96. Tomer, Y.; Wambach, J.; Knudsen, L.; Zhao, M.; Rodriguez, L.R.; Murthy, A.; White, F.V.; Venosa, A.; Katzen, J.; Ochs, M.; et al. The common ABCA3(E292V) variant disrupts AT2 cell quality control and increases susceptibility to lung injury and aberrant remodeling. Am. J. Physiol. Lung Cell Mol. Physiol. 2021, 321, L291–L307. [Google Scholar] [CrossRef]
  97. Dickens, J.A.; Rutherford, E.N.; Abreu, S.; Chambers, J.E.; Ellis, M.O.; van Schadewijk, A.; Hiemstra, P.S.; Marciniak, S.J. Novel insights into surfactant protein C trafficking revealed through the study of a pathogenic mutant. Eur. Respir. J. 2022, 59, 2100267. [Google Scholar] [CrossRef]
  98. van Moorsel, C.H.; Hoffman, T.W.; van Batenburg, A.A.; Klay, D.; van der Vis, J.J.; Grutters, J.C. Understanding Idiopathic Interstitial Pneumonia: A Gene-Based Review of Stressed Lungs. Biomed. Res. Int. 2015, 2015, 304186. [Google Scholar] [CrossRef] [Green Version]
  99. Legendre, M.; Butt, A.; Borie, R.; Debray, M.P.; Bouvry, D.; Filhol-Blin, E.; Desroziers, T.; Nau, V.; Copin, B.; Dastot-Le Moal, F.; et al. Functional assessment and phenotypic heterogeneity of SFTPA1 and SFTPA2 mutations in interstitial lung diseases and lung cancer. Eur. Respir. J. 2020, 56, 2002806. [Google Scholar] [CrossRef]
  100. Griese, M.; Lorenz, E.; Hengst, M.; Schams, A.; Wesselak, T.; Rauch, D.; Wittmann, T.; Kirchberger, V.; Escribano, A.; Schaible, T.; et al. Surfactant proteins in pediatric interstitial lung disease. Pediatr Res. 2016, 79, 34–41. [Google Scholar] [CrossRef] [Green Version]
  101. Salerno, T.; Peca, D.; Menchini, L.; Schiavino, A.; Boldrini, R.; Esposito, F.; Danhaive, O.; Cutrera, R. Surfactant Protein C-associated interstitial lung disease; three different phenotypes of the same SFTPC mutation. Ital. J. Pediatr. 2016, 42, 23. [Google Scholar] [CrossRef]
  102. van Moorsel, C.H.; Ten Klooster, L.; van Oosterhout, M.F.; de Jong, P.A.; Adams, H.; Wouter van Es, H.; Ruven, H.J.; van der Vis, J.J.; Grutters, J.C. SFTPA2 Mutations in Familial and Sporadic Idiopathic Interstitial Pneumonia. Am. J. Respir. Crit. Care Med. 2015, 192, 1249–1252. [Google Scholar] [CrossRef]
  103. Hayasaka, I.; Cho, K.; Akimoto, T.; Ikeda, M.; Uzuki, Y.; Yamada, M.; Nakata, K.; Furuta, I.; Ariga, T.; Minakami, H. Genetic basis for childhood interstitial lung disease among Japanese infants and children. Pediatr. Res. 2018, 83, 477–483. [Google Scholar] [CrossRef] [Green Version]
  104. Chen, J.; Nong, G.; Liu, X.; Ji, W.; Zhao, D.; Ma, H.; Wang, H.; Zheng, Y.; Shen, K. Genetic basis of surfactant dysfunction in Chinese children: A retrospective study. Pediatr. Pulmonol. 2019, 54, 1173–1181. [Google Scholar] [CrossRef]
  105. Klay, D.; Platenburg, M.; van Rijswijk, R.; Grutters, J.C.; van Moorsel, C.H.M. ABCA3 mutations in adult pulmonary fibrosis patients: A case series and review of literature. Curr. Opin. Pulm. Med. 2020, 26, 293–301. [Google Scholar] [CrossRef]
  106. Legendre, M.; Darde, X.; Ferreira, M.; Chantot-Bastaraud, S.; Campana, M.; Plantier, L.; Nathan, N.; Amselem, S.; Toutain, A.; Diot, P.; et al. The clinical course of interstitial lung disease in an adult patient with an ABCA3 homozygous complex allele under hydroxychloroquine and a review of the literature. Sarcoidosis Vasc. Diffuse Lung Dis. 2022, 39, e2022019. [Google Scholar] [CrossRef]
  107. Cho, J.G.; Thakkar, D.; Buchanan, P.; Graf, N.; Wheatley, J. ABCA3 deficiency from birth to adulthood presenting as paediatric interstitial lung disease. Respirol. Case Rep. 2020, 8, e00633. [Google Scholar] [CrossRef]
  108. Tang, X.; Li, H.; Liu, H.; Xu, H.; Yang, H.; Liu, J.; Zhao, S. Etiologic spectrum of interstitial lung diseases in Chinese children older than 2 years of age. Orphanet. J. Rare Dis. 2020, 15, 25. [Google Scholar] [CrossRef]
  109. Kröner, C.; Reu, S.; Teusch, V.; Schams, A.; Grimmelt, A.C.; Barker, M.; Brand, J.; Gappa, M.; Kitz, R.; Kramer, B.W.; et al. Genotype alone does not predict the clinical course of SFTPC deficiency in paediatric patients. Eur. Respir. J. 2015, 46, 197–206. [Google Scholar] [CrossRef] [Green Version]
  110. Chibbar, R.; Shih, F.; Baga, M.; Torlakovic, E.; Ramlall, K.; Skomro, R.; Cockcroft, D.W.; Lemire, E.G. Nonspecific interstitial pneumonia and usual interstitial pneumonia with mutation in surfactant protein C in familial pulmonary fibrosis. Mod. Pathol. 2004, 17, 973–980. [Google Scholar] [CrossRef] [Green Version]
  111. Klay, D.; Hoffman, T.W.; Harmsze, A.M.; Grutters, J.C.; van Moorsel, C.H.M. Systematic review of drug effects in humans and models with surfactant-processing disease. Eur. Respir. Rev. 2018, 27, 170135. [Google Scholar] [CrossRef]
  112. Griese, M.; Kappler, M.; Stehling, F.; Schulze, J.; Baden, W.; Koerner-Rettberg, C.; Carlens, J.; Prenzel, F.; Nährlich, L.; Thalmeier, A.; et al. Randomized controlled phase 2 trial of hydroxychloroquine in childhood interstitial lung disease. Orphanet. J. Rare Dis. 2022, 17, 289. [Google Scholar] [CrossRef]
  113. Cooney, A.L.; Wambach, J.A.; Sinn, P.L.; McCray, P.B., Jr. Gene Therapy Potential for Genetic Disorders of Surfactant Dysfunction. Front. Genome Ed. 2021, 3, 785829. [Google Scholar] [CrossRef]
  114. Kinting, S.; Höppner, S.; Schindlbeck, U.; Forstner, M.E.; Harfst, J.; Wittmann, T.; Griese, M. Functional rescue of misfolding ABCA3 mutations by small molecular correctors. Hum. Mol. Genet. 2018, 27, 943–953. [Google Scholar] [CrossRef] [Green Version]
  115. Eldridge, W.B.; Zhang, Q.; Faro, A.; Sweet, S.C.; Eghtesady, P.; Hamvas, A.; Cole, F.S.; Wambach, J.A. Outcomes of Lung Transplantation for Infants and Children with Genetic Disorders of Surfactant Metabolism. J. Pediatr. 2017, 184, 157–164.e2. [Google Scholar] [CrossRef] [Green Version]
  116. Armanios, M.Y.; Chen, J.J.; Cogan, J.D.; Alder, J.K.; Ingersoll, R.G.; Markin, C.; Lawson, W.E.; Xie, M.; Vulto, I.; Phillips, J.A., 3rd; et al. Telomerase mutations in families with idiopathic pulmonary fibrosis. N. Engl. J. Med. 2007, 356, 1317–1326. [Google Scholar] [CrossRef] [Green Version]
  117. Tsakiri, K.D.; Cronkhite, J.T.; Kuan, P.J.; Xing, C.; Raghu, G.; Weissler, J.C.; Rosenblatt, R.L.; Shay, J.W.; Garcia, C.K. Adult-onset pulmonary fibrosis caused by mutations in telomerase. Proc. Natl. Acad. Sci. USA 2007, 104, 7552–7557. [Google Scholar] [CrossRef] [Green Version]
  118. Vulliamy, T.J.; Walne, A.; Baskaradas, A.; Mason, P.J.; Marrone, A.; Dokal, I. Mutations in the reverse transcriptase component of telomerase (TERT) in patients with bone marrow failure. Blood Cells Mol. Dis. 2005, 34, 257–263. [Google Scholar] [CrossRef]
  119. Walne, A.J.; Dokal, I. Telomerase dysfunction and dyskeratosis congenita. Cytotechnology 2004, 45, 13–22. [Google Scholar] [CrossRef] [Green Version]
  120. Mason, P.J.; Bessler, M. The genetics of dyskeratosis congenita. Cancer Genet. 2011, 204, 635–645. [Google Scholar] [CrossRef] [Green Version]
  121. Bertuch, A.A. The molecular genetics of the telomere biology disorders. RNA Biol. 2016, 13, 696–706. [Google Scholar] [CrossRef] [PubMed]
  122. Vulliamy, T.J.; Marrone, A.; Knight, S.W.; Walne, A.; Mason, P.J.; Dokal, I. Mutations in dyskeratosis congenita: Their impact on telomere length and the diversity of clinical presentation. Blood 2006, 107, 2680–2685. [Google Scholar] [CrossRef] [PubMed]
  123. Savage, S.A.; Stewart, B.J.; Weksler, B.B.; Baerlocher, G.M.; Lansdorp, P.M.; Chanock, S.J.; Alter, B.P. Mutations in the reverse transcriptase component of telomerase (TERT) in patients with bone marrow failure. Blood Cells Mol. Dis. 2006, 37, 134–136. [Google Scholar] [CrossRef] [PubMed]
  124. Marrone, A.; Walne, A.; Tamary, H.; Masunari, Y.; Kirwan, M.; Beswick, R.; Vulliamy, T.; Dokal, I. Telomerase reverse-transcriptase homozygous mutations in autosomal recessive dyskeratosis congenita and Hoyeraal-Hreidarsson syndrome. Blood 2007, 110, 4198–4205. [Google Scholar] [CrossRef] [Green Version]
  125. Heiss, N.S.; Knight, S.W.; Vulliamy, T.J.; Klauck, S.M.; Wiemann, S.; Mason, P.J.; Poustka, A.; Dokal, I. X-linked dyskeratosis congenita is caused by mutations in a highly conserved gene with putative nucleolar functions. Nat. Genet. 1998, 19, 32–38. [Google Scholar] [CrossRef]
  126. Revy, P.; Kannengiesser, C.; Bertuch, A.A. Genetics of human telomere biology disorders. Nat. Rev. Genet. 2022. [Google Scholar] [CrossRef]
  127. Alder, J.K.; Armanios, M. Telomere-mediated lung disease. Physiol. Rev. 2022, 102, 1703–1720. [Google Scholar] [CrossRef]
  128. Borie, R.; Bouvry, D.; Cottin, V.; Gauvain, C.; Cazes, A.; Debray, M.P.; Cadranel, J.; Dieude, P.; Degot, T.; Dominique, S.; et al. Regulator of telomere length 1 (RTEL1) mutations are associated with heterogeneous pulmonary and extra-pulmonary phenotypes. Eur. Respir. J. 2019, 53, 1800508. [Google Scholar] [CrossRef]
  129. Benyelles, M.; O’Donohue, M.F.; Kermasson, L.; Lainey, E.; Borie, R.; Lagresle-Peyrou, C.; Nunes, H.; Cazelles, C.; Fourrage, C.; Ollivier, E.; et al. NHP2 deficiency impairs rRNA biogenesis and causes pulmonary fibrosis and Høyeraal-Hreidarsson syndrome. Hum. Mol. Genet. 2020, 29, 907–922. [Google Scholar] [CrossRef]
  130. Kannengiesser, C.; Manali, E.D.; Revy, P.; Callebaut, I.; Ba, I.; Borgel, A.; Oudin, C.; Haritou, A.; Kolilekas, L.; Malagari, K.; et al. First heterozygous NOP10 mutation in familial pulmonary fibrosis. Eur. Respir. J. 2020, 55, 1902465. [Google Scholar] [CrossRef]
  131. Kropski, J.A.; Reiss, S.; Markin, C.; Brown, K.K.; Schwartz, D.A.; Schwarz, M.I.; Loyd, J.E.; Phillips, J.A., 3rd; Blackwell, T.S.; Cogan, J.D. Rare Genetic Variants in PARN Are Associated with Pulmonary Fibrosis in Families. Am. J. Respir. Crit. Care Med. 2017, 196, 1481–1484. [Google Scholar] [CrossRef]
  132. Stuart, B.D.; Choi, J.; Zaidi, S.; Xing, C.; Holohan, B.; Chen, R.; Choi, M.; Dharwadkar, P.; Torres, F.; Girod, C.E.; et al. Exome sequencing links mutations in PARN and RTEL1 with familial pulmonary fibrosis and telomere shortening. Nat. Genet. 2015, 47, 512–517. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Gable, D.L.; Gaysinskaya, V.; Atik, C.C.; Talbot, C.C., Jr.; Kang, B.; Stanley, S.E.; Pugh, E.W.; Amat-Codina, N.; Schenk, K.M.; Arcasoy, M.O.; et al. ZCCHC8, the nuclear exosome targeting component, is mutated in familial pulmonary fibrosis and is required for telomerase RNA maturation. Genes Dev. 2019, 33, 1381–1396. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Kelich, J.; Aramburu, T.; van der Vis, J.J.; Showe, L.; Kossenkov, A.; van der Smagt, J.; Massink, M.; Schoemaker, A.; Hennekam, E.; Veltkamp, M.; et al. Telomere dysfunction implicates POT1 in patients with idiopathic pulmonary fibrosis. J. Exp. Med. 2022, 219, e20211681. [Google Scholar] [CrossRef]
  135. Stanley, S.E.; Gable, D.L.; Wagner, C.L.; Carlile, T.M.; Hanumanthu, V.S.; Podlevsky, J.D.; Khalil, S.E.; DeZern, A.E.; Rojas-Duran, M.F.; Applegate, C.D.; et al. Loss-of-function mutations in the RNA biogenesis factor NAF1 predispose to pulmonary fibrosis-emphysema. Sci. Transl. Med. 2016, 8, 351ra107. [Google Scholar] [CrossRef] [Green Version]
  136. Gaysinskaya, V.; Stanley, S.E.; Adam, S.; Armanios, M. Synonymous Mutation in DKC1 Causes Telomerase RNA Insufficiency Manifesting as Familial Pulmonary Fibrosis. Chest 2020, 158, 2449–2457. [Google Scholar] [CrossRef]
  137. Alder, J.K.; Stanley, S.E.; Wagner, C.L.; Hamilton, M.; Hanumanthu, V.S.; Armanios, M. Exome sequencing identifies mutant TINF2 in a family with pulmonary fibrosis. Chest 2015, 147, 1361–1368. [Google Scholar] [CrossRef] [Green Version]
  138. van der Vis, J.J.; van der Smagt, J.J.; Hennekam, F.A.M.; Grutters, J.C.; van Moorsel, C.H.M. Pulmonary Fibrosis and a TERT Founder Mutation With a Latency Period of 300 Years. Chest 2020, 158, 612–619. [Google Scholar] [CrossRef]
  139. Sharma, R.; Sahoo, S.S.; Honda, M.; Granger, S.L.; Goodings, C.; Sanchez, L.; Künstner, A.; Busch, H.; Beier, F.; Pruett-Miller, S.M.; et al. Gain-of-function mutations in RPA1 cause a syndrome with short telomeres and somatic genetic rescue. Blood 2022, 139, 1039–1051. [Google Scholar] [CrossRef]
  140. M’Kacher, R.; Jaillet, M.; Colicchio, B.; Vasarmidi, E.; Mailleux, A.; Dieterlen, A.; Kannengiesser, C.; Borie, C.; Oudrhiri, N.; Junker, S.; et al. Lung Fibroblasts from Idiopathic Pulmonary Fibrosis Patients Harbor Short and Unstable Telomeres Leading to Chromosomal Instability. Biomedicines 2022, 10, 310. [Google Scholar] [CrossRef]
  141. Schratz, K.E.; Haley, L.; Danoff, S.K.; Blackford, A.L.; DeZern, A.E.; Gocke, C.D.; Duffield, A.S.; Armanios, M. Cancer spectrum and outcomes in the Mendelian short telomere syndromes. Blood 2020, 135, 1946–1956. [Google Scholar] [CrossRef]
  142. Schratz, K.E. Extrahematopoietic manifestations of the short telomere syndromes. Hematol. Am. Soc. Hematol. Educ. Program 2020, 2020, 115–122. [Google Scholar] [CrossRef]
  143. Wagner, C.L.; Hanumanthu, V.S.; Talbot, C.C., Jr.; Abraham, R.S.; Hamm, D.; Gable, D.L.; Kanakry, C.G.; Applegate, C.D.; Siliciano, J.; Jackson, J.B.; et al. Short telomere syndromes cause a primary T cell immunodeficiency. J. Clin. Investig. 2018, 128, 5222–5234. [Google Scholar] [CrossRef] [Green Version]
  144. Tomos, I.; Karakatsani, A.; Manali, E.D.; Kottaridi, C.; Spathis, A.; Argentos, S.; Papiris, S.A. Telomere length across different UIP fibrotic-Interstitial Lung Diseases: A prospective Greek case-control study. Pulmonology 2022, 28, 254–261. [Google Scholar] [CrossRef]
  145. Alder, J.K.; Chen, J.J.; Lancaster, L.; Danoff, S.; Su, S.C.; Cogan, J.D.; Vulto, I.; Xie, M.; Qi, X.; Tuder, R.M.; et al. Short telomeres are a risk factor for idiopathic pulmonary fibrosis. Proc. Natl. Acad. Sci. USA 2008, 105, 13051–13056. [Google Scholar] [CrossRef] [Green Version]
  146. Alder, J.K.; Hanumanthu, V.S.; Strong, M.A.; DeZern, A.E.; Stanley, S.E.; Takemoto, C.M.; Danilova, L.; Applegate, C.D.; Bolton, S.G.; Mohr, D.W.; et al. Diagnostic utility of telomere length testing in a hospital-based setting. Proc. Natl. Acad. Sci. USA 2018, 115, E2358–E2365. [Google Scholar] [CrossRef] [Green Version]
  147. Gansner, J.M.; Rosas, I.O.; Ebert, B.L. Pulmonary fibrosis, bone marrow failure, and telomerase mutation. N. Engl. J. Med. 2012, 366, 1551–1553. [Google Scholar] [CrossRef]
  148. Armanios, M.; Alder, J.K.; Parry, E.M.; Karim, B.; Strong, M.A.; Greider, C.W. Short telomeres are sufficient to cause the degenerative defects associated with aging. Am. J. Hum. Genet. 2009, 85, 823–832. [Google Scholar] [CrossRef] [Green Version]
  149. Hong, X.; Wang, L.; Zhang, K.; Liu, J.; Liu, J.P. Molecular Mechanisms of Alveolar Epithelial Stem Cell Senescence and Senescence-Associated Differentiation Disorders in Pulmonary Fibrosis. Cells 2022, 11, 877. [Google Scholar] [CrossRef]
  150. Roake, C.M.; Artandi, S.E. Regulation of human telomerase in homeostasis and disease. Nat. Rev. Mol. Cell Biol. 2020, 21, 384–397. [Google Scholar] [CrossRef]
  151. Tsang, A.R.; Wyatt, H.D.; Ting, N.S.; Beattie, T.L. hTERT mutations associated with idiopathic pulmonary fibrosis affect telomerase activity, telomere length, and cell growth by distinct mechanisms. Aging Cell 2012, 11, 482–490. [Google Scholar] [CrossRef]
  152. Armanios, M.; Blackburn, E.H. The telomere syndromes. Nat. Rev. Genet. 2012, 13, 693–704. [Google Scholar] [CrossRef]
  153. Molina-Molina, M.; Borie, R. Clinical implications of telomere dysfunction in lung fibrosis. Curr. Opin. Pulm. Med. 2018, 24, 440–444. [Google Scholar] [CrossRef]
  154. Salehian, S.; Semple, T.; Pabary, R. Childhood interstitial lung disease: Short lessons from telomeres. Thorax 2021, 76, 1250–1252. [Google Scholar] [CrossRef]
  155. Alder, J.K.; Cogan, J.D.; Brown, A.F.; Anderson, C.J.; Lawson, W.E.; Lansdorp, P.M.; Phillips, J.A., 3rd; Loyd, J.E.; Chen, J.J.; Armanios, M. Ancestral mutation in telomerase causes defects in repeat addition processivity and manifests as familial pulmonary fibrosis. PLoS Genet. 2011, 7, e1001352. [Google Scholar] [CrossRef] [Green Version]
  156. Silhan, L.L.; Shah, P.D.; Chambers, D.C.; Snyder, L.D.; Riise, G.C.; Wagner, C.L.; Hellström-Lindberg, E.; Orens, J.B.; Mewton, J.F.; Danoff, S.K.; et al. Lung transplantation in telomerase mutation carriers with pulmonary fibrosis. Eur. Respir. J. 2014, 44, 178–187. [Google Scholar] [CrossRef] [Green Version]
  157. Borie, R.; Kannengiesser, C.; Hirschi, S.; Le Pavec, J.; Mal, H.; Bergot, E.; Jouneau, S.; Naccache, J.M.; Revy, P.; Boutboul, D.; et al. Severe hematologic complications after lung transplantation in patients with telomerase complex mutations. J. Heart. Lung Transpl. 2015, 34, 538–546. [Google Scholar] [CrossRef]
  158. Phillips-Houlbracq, M.; Mal, H.; Cottin, V.; Gauvain, C.; Beier, F.; Sicre de Fontbrune, F.; Sidali, S.; Mornex, J.F.; Hirschi, S.; Roux, A.; et al. Determinants of survival after lung transplantation in telomerase-related gene mutation carriers: A retrospective cohort. Am. J. Transpl. 2022, 22, 1236–1244. [Google Scholar] [CrossRef]
  159. Swaminathan, A.C.; Neely, M.L.; Frankel, C.W.; Kelly, F.L.; Petrovski, S.; Durheim, M.T.; Bush, E.; Snyder, L.; Goldstein, D.B.; Todd, J.L.; et al. Lung Transplant Outcomes in Patients With Pulmonary Fibrosis With Telomere-Related Gene Variants. Chest 2019, 156, 477–485. [Google Scholar] [CrossRef]
  160. Choi, B.; Messika, J.; Courtwright, A.; Mornex, J.F.; Hirschi, S.; Roux, A.; Le Pavec, J.; Quêtant, S.; Froidure, A.; Lazor, R.; et al. Airway complications in lung transplant recipients with telomere-related interstitial lung disease. Clin. Transpl. 2022, 36, e14552. [Google Scholar] [CrossRef]
  161. George, G.; Rosas, I.O.; Cui, Y.; McKane, C.; Hunninghake, G.M.; Camp, P.C.; Raby, B.A.; Goldberg, H.J.; El-Chemaly, S. Short telomeres, telomeropathy, and subclinical extrapulmonary organ damage in patients with interstitial lung disease. Chest 2015, 147, 1549–1557. [Google Scholar] [CrossRef]
  162. Le Pavec, J.; Dauriat, G.; Gazengel, P.; Dolidon, S.; Hanna, A.; Feuillet, S.; Pradere, P.; Crutu, A.; Florea, V.; Boulate, D.; et al. Lung transplantation for idiopathic pulmonary fibrosis. Presse Med. 2020, 49, 104026. [Google Scholar] [CrossRef]
  163. Popescu, I.; Mannem, H.; Winters, S.A.; Hoji, A.; Silveira, F.; McNally, E.; Pipeling, M.R.; Lendermon, E.A.; Morrell, M.R.; Pilewski, J.M.; et al. Impaired Cytomegalovirus Immunity in Idiopathic Pulmonary Fibrosis Lung Transplant Recipients with Short Telomeres. Am. J. Respir. Crit. Care Med. 2019, 199, 362–376. [Google Scholar] [CrossRef]
  164. Borie, R.; Kannengiesser, C.; Sicre de Fontbrune, F.; Boutboul, D.; Tabeze, L.; Brunet-Possenti, F.; Lainey, E.; Debray, M.P.; Cazes, A.; Crestani, B. Pneumocystosis revealing immunodeficiency secondary to TERC mutation. Eur. Respir. J. 2017, 50, 1701443. [Google Scholar] [CrossRef] [Green Version]
  165. Raghu, G.; Anstrom, K.J.; King, T.E., Jr.; Lasky, J.A.; Martinez, F.J. Prednisone, azathioprine, and N-acetylcysteine for pulmonary fibrosis. N. Engl. J. Med. 2012, 366, 1968–1977. [Google Scholar] [CrossRef]
  166. Newton, C.A.; Zhang, D.; Oldham, J.M.; Kozlitina, J.; Ma, S.F.; Martinez, F.J.; Raghu, G.; Noth, I.; Garcia, C.K. Telomere Length and Use of Immunosuppressive Medications in Idiopathic Pulmonary Fibrosis. Am. J. Respir. Crit. Care Med. 2019, 200, 336–347. [Google Scholar] [CrossRef]
  167. Reilly, C.R.; Myllymäki, M.; Redd, R.; Padmanaban, S.; Karunakaran, D.; Tesmer, V.; Tsai, F.D.; Gibson, C.J.; Rana, H.Q.; Zhong, L.; et al. The clinical and functional effects of TERT variants in myelodysplastic syndrome. Blood 2021, 138, 898–911. [Google Scholar] [CrossRef]
  168. Parry, E.M.; Alder, J.K.; Qi, X.; Chen, J.J.-L.; Armanios, M. Syndrome complex of bone marrow failure and pulmonary fibrosis predicts germline defects in telomerase. Blood 2011, 117, 5607–5611, Erratum in Blood 2016, 127, 1837. [Google Scholar] [CrossRef] [Green Version]
  169. Papiris, S.A.; Tsirigotis, P.; Kannengiesser, C.; Kolilekas, L.; Gkirkas, K.; Papaioannou, A.I.; Revy, P.; Giouleka, P.; Papadaki, G.; Kagouridis, K.; et al. Myelodysplastic syndromes and idiopathic pulmonary fibrosis: A dangerous liaison. Respir. Res. 2019, 20, 182. [Google Scholar] [CrossRef] [Green Version]
  170. Diaz de Leon, A.; Cronkhite, J.T.; Katzenstein, A.L.; Godwin, J.D.; Raghu, G.; Glazer, C.S.; Rosenblatt, R.L.; Girod, C.E.; Garrity, E.R.; Xing, C.; et al. Telomere lengths, pulmonary fibrosis and telomerase (TERT) mutations. PLoS ONE 2010, 5, e10680. [Google Scholar] [CrossRef] [Green Version]
  171. Schratz, K.E.; Gaysinskaya, V.; Cosner, Z.L.; DeBoy, E.A.; Xiang, Z.; Kasch-Semenza, L.; Florea, L.; Shah, P.D.; Armanios, M. Somatic reversion impacts myelodysplastic syndromes and acute myeloid leukemia evolution in the short telomere disorders. J. Clin. Investig. 2021, 131, e147598. [Google Scholar] [CrossRef]
  172. Calado, R.T.; Regal, J.A.; Kleiner, D.E.; Schrump, D.S.; Peterson, N.R.; Pons, V.; Chanock, S.J.; Lansdorp, P.M.; Young, N.S. A spectrum of severe familial liver disorders associate with telomerase mutations. PLoS ONE 2009, 4, e7926. [Google Scholar] [CrossRef] [Green Version]
  173. Calado, R.T.; Brudno, J.; Mehta, P.; Kovacs, J.J.; Wu, C.; Zago, M.A.; Chanock, S.J.; Boyer, T.D.; Young, N.S. Constitutional telomerase mutations are genetic risk factors for cirrhosis. Hepatology 2011, 53, 1600–1607. [Google Scholar] [CrossRef] [Green Version]
  174. Nault, J.C.; Ningarhari, M.; Rebouissou, S.; Zucman-Rossi, J. The role of telomeres and telomerase in cirrhosis and liver cancer. Nat. Rev. Gastroenterol. Hepatol. 2019, 16, 544–558. [Google Scholar] [CrossRef]
  175. Rudolph, K.L.; Chang, S.; Millard, M.; Schreiber-Agus, N.; DePinho, R.A. Inhibition of experimental liver cirrhosis in mice by telomerase gene delivery. Science 2000, 287, 1253–1258. [Google Scholar] [CrossRef]
  176. Gorgy, A.I.; Jonassaint, N.L.; Stanley, S.E.; Koteish, A.; DeZern, A.E.; Walter, J.E.; Sopha, S.C.; Hamilton, J.P.; Hoover-Fong, J.; Chen, A.R.; et al. Hepatopulmonary syndrome is a frequent cause of dyspnea in the short telomere disorders. Chest 2015, 148, 1019–1026. [Google Scholar] [CrossRef] [Green Version]
  177. Hoffman, T.W.; van der Vis, J.J.; Biesma, D.H.; Grutters, J.C.; van Moorsel, C.H.M. Extrapulmonary manifestations of a telomere syndrome in patients with idiopathic pulmonary fibrosis are associated with decreased survival. Respirology 2022, 27, 959–965. [Google Scholar] [CrossRef]
  178. Raghu, G.; Rochwerg, B.; Zhang, Y.; Garcia, C.A.; Azuma, A.; Behr, J.; Brozek, J.L.; Collard, H.R.; Cunningham, W.; Homma, S.; et al. An Official ATS/ERS/JRS/ALAT Clinical Practice Guideline: Treatment of Idiopathic Pulmonary Fibrosis. An Update of the 2011 Clinical Practice Guideline. Am. J. Respir. Crit. Care Med. 2015, 192, e3–e19. [Google Scholar] [CrossRef]
  179. Justet, A.; Thabut, G.; Manali, E.; Molina Molina, M.; Kannengiesser, C.; Cadranel, J.; Cottin, V.; Gondouin, A.; Nunes, H.; Magois, E.; et al. Safety and efficacy of pirfenidone in patients carrying telomerase complex mutation. Eur. Respir. J. 2018, 51, 1701875. [Google Scholar] [CrossRef] [Green Version]
  180. Justet, A.; Klay, D.; Porcher, R.; Cottin, V.; Ahmad, K.; Molina Molina, M.; Nunes, H.; Reynaud-Gaubert, M.; Naccache, J.M.; Manali, E.; et al. Safety and efficacy of pirfenidone and nintedanib in patients with idiopathic pulmonary fibrosis and carrying a telomere-related gene mutation. Eur. Respir. J. 2021, 57, 2003198. [Google Scholar] [CrossRef]
  181. Townsley, D.M.; Dumitriu, B.; Liu, D.; Biancotto, A.; Weinstein, B.; Chen, C.; Hardy, N.; Mihalek, A.D.; Lingala, S.; Kim, Y.J.; et al. Danazol Treatment for Telomere Diseases. N. Engl. J. Med. 2016, 374, 1922–1931. [Google Scholar] [CrossRef] [PubMed]
  182. Mackintosh, J.A.; Pietsch, M.; Lutzky, V.; Enever, D.; Bancroft, S.; Apte, S.H.; Tan, M.; Yerkovich, S.T.; Dickinson, J.L.; Pickett, H.A.; et al. TELO-SCOPE study: A randomised, double-blind, placebo-controlled, phase 2 trial of danazol for short telomere related pulmonary fibrosis. BMJ Open Respir. Res. 2021, 8, e001127. [Google Scholar] [CrossRef]
  183. Basil, M.C.; Katzen, J.; Engler, A.E.; Guo, M.; Herriges, M.J.; Kathiriya, J.J.; Windmueller, R.; Ysasi, A.B.; Zacharias, W.J.; Chapman, H.A.; et al. The Cellular and Physiological Basis for Lung Repair and Regeneration: Past, Present, and Future. Cell Stem Cell 2020, 26, 482–502. [Google Scholar] [CrossRef]
  184. Liu, Y.; Jesus, A.A.; Marrero, B.; Yang, D.; Ramsey, S.E.; Sanchez, G.A.M.; Tenbrock, K.; Wittkowski, H.; Jones, O.Y.; Kuehn, H.S.; et al. Activated STING in a vascular and pulmonary syndrome. N. Engl. J. Med. 2014, 371, 507–518. [Google Scholar] [CrossRef] [Green Version]
  185. David, C.; Frémond, M.L. Lung Inflammation in STING-Associated Vasculopathy with Onset in Infancy (SAVI). Cells 2022, 11, 318. [Google Scholar] [CrossRef]
  186. Picard, C.; Thouvenin, G.; Kannengiesser, C.; Dubus, J.C.; Jeremiah, N.; Rieux-Laucat, F.; Crestani, B.; Belot, A.; Thivolet-Béjui, F.; Secq, V.; et al. Severe Pulmonary Fibrosis as the First Manifestation of Interferonopathy (TMEM173 Mutation). Chest 2016, 150, e65–e71. [Google Scholar] [CrossRef] [Green Version]
  187. Staels, F.; Betrains, A.; Doubel, P.; Willemsen, M.; Cleemput, V.; Vanderschueren, S.; Corveleyn, A.; Meyts, I.; Sprangers, B.; Crow, Y.J.; et al. Adult-Onset ANCA-Associated Vasculitis in SAVI: Extension of the Phenotypic Spectrum, Case Report and Review of the Literature. Front. Immunol. 2020, 11, 575219. [Google Scholar] [CrossRef]
  188. Watkin, L.B.; Jessen, B.; Wiszniewski, W.; Vece, T.J.; Jan, M.; Sha, Y.; Thamsen, M.; Santos-Cortez, R.L.; Lee, K.; Gambin, T.; et al. COPA mutations impair ER-Golgi transport and cause hereditary autoimmune-mediated lung disease and arthritis. Nat. Genet. 2015, 47, 654–660. [Google Scholar] [CrossRef] [Green Version]
  189. Steiner, A.; Hrovat-Schaale, K.; Prigione, I.; Yu, C.H.; Laohamonthonkul, P.; Harapas, C.R.; Low, R.R.J.; De Nardo, D.; Dagley, L.F.; Mlodzianoski, M.J.; et al. Deficiency in coatomer complex I causes aberrant activation of STING signalling. Nat. Commun. 2022, 13, 2321. [Google Scholar] [CrossRef]
  190. Lepelley, A.; Martin-Niclós, M.J.; Le Bihan, M.; Marsh, J.A.; Uggenti, C.; Rice, G.I.; Bondet, V.; Duffy, D.; Hertzog, J.; Rehwinkel, J.; et al. Mutations in COPA lead to abnormal trafficking of STING to the Golgi and interferon signaling. J. Exp. Med. 2020, 217, e20200600. [Google Scholar] [CrossRef]
  191. Vece, T.J.; Watkin, L.B.; Nicholas, S.; Canter, D.; Braun, M.C.; Guillerman, R.P.; Eldin, K.W.; Bertolet, G.; McKinley, S.; de Guzman, M.; et al. Copa Syndrome: A Novel Autosomal Dominant Immune Dysregulatory Disease. J. Clin. Immunol. 2016, 36, 377–387. [Google Scholar] [CrossRef]
  192. Volpi, S.; Tsui, J.; Mariani, M.; Pastorino, C.; Caorsi, R.; Sacco, O.; Ravelli, A.; Shum, A.K.; Gattorno, M.; Picco, P. Type I interferon pathway activation in COPA syndrome. Clin. Immunol. 2018, 187, 33–36. [Google Scholar] [CrossRef]
  193. Yokoyama, T.; Gochuico, B.R. Hermansky-Pudlak syndrome pulmonary fibrosis: A rare inherited interstitial lung disease. Eur. Respir. Rev. 2021, 30, 200193. [Google Scholar] [CrossRef]
  194. Hermansky, F.; Pudlak, P. Albinism associated with hemorrhagic diathesis and unusual pigmented reticular cells in the bone marrow: Report of two cases with histochemical studies. Blood 1959, 14, 162–169. [Google Scholar] [CrossRef] [Green Version]
  195. Oh, J.; Bailin, T.; Fukai, K.; Feng, G.H.; Ho, L.; Mao, J.I.; Frenk, E.; Tamura, N.; Spritz, R.A. Positional cloning of a gene for Hermansky-Pudlak syndrome, a disorder of cytoplasmic organelles. Nat. Genet. 1996, 14, 300–306. [Google Scholar] [CrossRef]
  196. Dell’Angelica, E.C.; Shotelersuk, V.; Aguilar, R.C.; Gahl, W.A.; Bonifacino, J.S. Altered trafficking of lysosomal proteins in Hermansky-Pudlak syndrome due to mutations in the beta 3A subunit of the AP-3 adaptor. Mol. Cell 1999, 3, 11–21. [Google Scholar] [CrossRef]
  197. Anderson, P.D.; Huizing, M.; Claassen, D.A.; White, J.; Gahl, W.A. Hermansky-Pudlak syndrome type 4 (HPS-4): Clinical and molecular characteristics. Hum. Genet. 2003, 113, 10–17. [Google Scholar] [CrossRef] [PubMed]
  198. Ammann, S.; Schulz, A.; Krägeloh-Mann, I.; Dieckmann, N.M.; Niethammer, K.; Fuchs, S.; Eckl, K.M.; Plank, R.; Werner, R.; Altmüller, J.; et al. Mutations in AP3D1 associated with immunodeficiency and seizures define a new type of Hermansky-Pudlak syndrome. Blood 2016, 127, 997–1006. [Google Scholar] [CrossRef] [Green Version]
  199. Huizing, M.; Malicdan, M.C.V.; Wang, J.A.; Pri-Chen, H.; Hess, R.A.; Fischer, R.; O’Brien, K.J.; Merideth, M.A.; Gahl, W.A.; Gochuico, B.R. Hermansky-Pudlak syndrome: Mutation update. Hum. Mutat. 2020, 41, 543–580. [Google Scholar] [CrossRef]
  200. Hengst, M.; Naehrlich, L.; Mahavadi, P.; Grosse-Onnebrink, J.; Terheggen-Lagro, S.; Skanke, L.H.; Schuch, L.A.; Brasch, F.; Guenther, A.; Reu, S.; et al. Hermansky-Pudlak syndrome type 2 manifests with fibrosing lung disease early in childhood. Orphanet J. Rare Dis. 2018, 13, 42. [Google Scholar] [CrossRef]
  201. Gochuico, B.R.; Huizing, M.; Golas, G.A.; Scher, C.D.; Tsokos, M.; Denver, S.D.; Frei-Jones, M.J.; Gahl, W.A. Interstitial lung disease and pulmonary fibrosis in Hermansky-Pudlak syndrome type 2, an adaptor protein-3 complex disease. Mol. Med. 2012, 18, 56–64. [Google Scholar] [CrossRef] [PubMed]
  202. Vicary, G.W.; Vergne, Y.; Santiago-Cornier, A.; Young, L.R.; Roman, J. Pulmonary Fibrosis in Hermansky-Pudlak Syndrome. Ann. Am. Thorac. Soc. 2016, 13, 1839–1846. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Rouhani, F.N.; Brantly, M.L.; Markello, T.C.; Helip-Wooley, A.; O’Brien, K.; Hess, R.; Huizing, M.; Gahl, W.A.; Gochuico, B.R. Alveolar macrophage dysregulation in Hermansky-Pudlak syndrome type 1. Am. J. Respir. Crit. Care Med. 2009, 180, 1114–1121. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Imani, J.; Bodine, S.P.M.; Lamattina, A.M.; Ma, D.D.; Shrestha, S.; Maynard, D.M.; Bishop, K.; Nwokeji, A.; Malicdan, M.C.V.; Testa, L.C.; et al. Dysregulated myosin in Hermansky-Pudlak syndrome lung fibroblasts is associated with increased cell motility. Respir. Res. 2022, 23, 167. [Google Scholar] [CrossRef]
  205. Lyerla, T.A.; Rusiniak, M.E.; Borchers, M.; Jahreis, G.; Tan, J.; Ohtake, P.; Novak, E.K.; Swank, R.T. Aberrant lung structure, composition, and function in a murine model of Hermansky-Pudlak syndrome. Am. J. Physiol. Lung Cell Mol. Physiol. 2003, 285, L643–L653. [Google Scholar] [CrossRef] [Green Version]
  206. Chiang, P.W.; Oiso, N.; Gautam, R.; Suzuki, T.; Swank, R.T.; Spritz, R.A. The Hermansky-Pudlak syndrome 1 (HPS1) and HPS4 proteins are components of two complexes, BLOC-3 and BLOC-4, involved in the biogenesis of lysosome-related organelles. J. Biol. Chem. 2003, 278, 20332–20337. [Google Scholar] [CrossRef] [Green Version]
  207. Atochina-Vasserman, E.N.; Bates, S.R.; Zhang, P.; Abramova, H.; Zhang, Z.; Gonzales, L.; Tao, J.Q.; Gochuico, B.R.; Gahl, W.; Guo, C.J.; et al. Early alveolar epithelial dysfunction promotes lung inflammation in a mouse model of Hermansky-Pudlak syndrome. Am. J. Respir. Crit. Care Med. 2011, 184, 449–458. [Google Scholar] [CrossRef] [Green Version]
  208. Young, L.R.; Gulleman, P.M.; Bridges, J.P.; Weaver, T.E.; Deutsch, G.H.; Blackwell, T.S.; McCormack, F.X. The alveolar epithelium determines susceptibility to lung fibrosis in Hermansky-Pudlak syndrome. Am. J. Respir. Crit. Care Med. 2012, 186, 1014–1024. [Google Scholar] [CrossRef] [Green Version]
  209. Yanagihara, T.; Sato, S.; Upagupta, C.; Kolb, M. What have we learned from basic science studies on idiopathic pulmonary fibrosis? Eur. Respir. Rev. 2019, 28, 190029. [Google Scholar] [CrossRef] [Green Version]
  210. O’Brien, K.J.; Introne, W.J.; Akal, O.; Akal, T.; Barbu, A.; McGowan, M.P.; Merideth, M.A.; Seward, S.L., Jr.; Gahl, W.A.; Gochuico, B.R. Prolonged treatment with open-label pirfenidone in Hermansky-Pudlak syndrome pulmonary fibrosis. Mol. Genet. Metab. 2018, 125, 168–173. [Google Scholar] [CrossRef]
  211. Benvenuto, L.; Qayum, S.; Kim, H.; Robbins, H.; Shah, L.; Dimango, A.; Magda, G.; Grewal, H.; Lemaitre, P.; Stanifer, B.P.; et al. Lung Transplantation for Pulmonary Fibrosis Associated With Hermansky-Pudlak Syndrome. A Single-center Experience. Transpl. Direct 2022, 8, e1303. [Google Scholar] [CrossRef] [PubMed]
  212. Cetin Gedik, K.; Lamot, L.; Romano, M.; Demirkaya, E.; Piskin, D.; Torreggiani, S.; Adang, L.A.; Armangue, T.; Barchus, K.; Cordova, D.R.; et al. The 2021 European Alliance of Associations for Rheumatology/American College of Rheumatology Points to Consider for Diagnosis and Management of Autoinflammatory Type I Interferonopathies: CANDLE/PRAAS, SAVI, and AGS. Arthritis Rheumatol. 2022, 74, 735–751. [Google Scholar] [CrossRef] [PubMed]
  213. Frémond, M.L.; Legendre, M.; Fayon, M.; Clement, A.; Filhol-Blin, E.; Richard, N.; Berdah, L.; Roullaud, S.; Rice, G.I.; Bondet, V.; et al. Use of ruxolitinib in COPA syndrome manifesting as life-threatening alveolar haemorrhage. Thorax 2020, 75, 92–95. [Google Scholar] [CrossRef] [PubMed]
  214. Griese, M.; Seidl, E.; Hengst, M.; Reu, S.; Rock, H.; Anthony, G.; Kiper, N.; Emiralioğlu, N.; Snijders, D.; Goldbeck, L.; et al. International management platform for children’s interstitial lung disease (chILD-EU). Thorax 2018, 73, 231–239. [Google Scholar] [CrossRef] [Green Version]
  215. Schwartz, D.A. Idiopathic pulmonary fibrosis is a complex genetic disorder. Trans. Am. Clin. Climatol. Assoc. 2016, 127, 34–45. [Google Scholar]
  216. Fingerlin, T.E.; Murphy, E.; Zhang, W.; Peljto, A.L.; Brown, K.K.; Steele, M.P.; Loyd, J.E.; Cosgrove, G.P.; Lynch, D.; Groshong, S.; et al. Genome-wide association study identifies multiple susceptibility loci for pulmonary fibrosis. Nat. Genet. 2013, 45, 613–620. [Google Scholar] [CrossRef] [Green Version]
  217. Korthagen, N.M.; van Moorsel, C.H.; Barlo, N.P.; Kazemier, K.M.; Ruven, H.J.; Grutters, J.C. Association between variations in cell cycle genes and idiopathic pulmonary fibrosis. PLoS ONE 2012, 7, e30442. [Google Scholar] [CrossRef]
  218. Pulkkinen, V.; Bruce, S.; Rintahaka, J.; Hodgson, U.; Laitinen, T.; Alenius, H.; Kinnula, V.L.; Myllärniemi, M.; Matikainen, S.; Kere, J. ELMOD2, a candidate gene for idiopathic pulmonary fibrosis, regulates antiviral responses. FASEB J. 2010, 24, 1167–1177. [Google Scholar] [CrossRef]
  219. Ahn, M.H.; Park, B.L.; Lee, S.H.; Park, S.W.; Park, J.S.; Kim, D.J.; Jang, A.S.; Park, J.S.; Shin, H.K.; Uh, S.T.; et al. A promoter SNP rs4073T>A in the common allele of the interleukin 8 gene is associated with the development of idiopathic pulmonary fibrosis via the IL-8 protein enhancing mode. Respir. Res. 2011, 12, 73. [Google Scholar] [CrossRef] [Green Version]
  220. Noth, I.; Zhang, Y.; Ma, S.F.; Flores, C.; Barber, M.; Huang, Y.; Broderick, S.M.; Wade, M.S.; Hysi, P.; Scuirba, J.; et al. Genetic variants associated with idiopathic pulmonary fibrosis susceptibility and mortality: A genome-wide association study. Lancet Respir. Med. 2013, 1, 309–317. [Google Scholar] [CrossRef] [Green Version]
  221. O’Dwyer, D.N.; Armstrong, M.E.; Trujillo, G.; Cooke, G.; Keane, M.P.; Fallon, P.G.; Simpson, A.J.; Millar, A.B.; McGrath, E.E.; Whyte, M.K.; et al. The Toll-like receptor 3 L412F polymorphism and disease progression in idiopathic pulmonary fibrosis. Am. J. Respir. Crit. Care Med. 2013, 188, 1442–1450. [Google Scholar] [CrossRef]
  222. Zhang, D.; Povysil, G.; Kobeissy, P.H.; Li, Q.; Wang, B.; Amelotte, M.; Jaouadi, H.; Newton, C.A.; Maher, T.M.; Molyneaux, P.L.; et al. Rare and Common Variants in KIF15 Contribute to Genetic Risk of Idiopathic Pulmonary Fibrosis. Am. J. Respir. Crit. Care Med. 2022, 206, 56–69. [Google Scholar] [CrossRef]
  223. Seibold, M.A.; Wise, A.L.; Speer, M.C.; Steele, M.P.; Brown, K.K.; Loyd, J.E.; Fingerlin, T.E.; Zhang, W.; Gudmundsson, G.; Groshong, S.D.; et al. A common MUC5B promoter polymorphism and pulmonary fibrosis. N. Engl. J. Med. 2011, 364, 1503–1512. [Google Scholar] [CrossRef] [Green Version]
  224. Horimasu, Y.; Ohshimo, S.; Bonella, F.; Tanaka, S.; Ishikawa, N.; Hattori, N.; Kohno, N.; Guzman, J.; Costabel, U. MUC5B promoter polymorphism in Japanese patients with idiopathic pulmonary fibrosis. Respirology 2015, 20, 439–444. [Google Scholar] [CrossRef]
  225. Peljto, A.L.; Selman, M.; Kim, D.S.; Murphy, E.; Tucker, L.; Pardo, A.; Lee, J.S.; Ji, W.; Schwarz, M.I.; Yang, I.V.; et al. The MUC5B promoter polymorphism is associated with idiopathic pulmonary fibrosis in a Mexican cohort but is rare among Asian ancestries. Chest 2015, 147, 460–464. [Google Scholar] [CrossRef] [Green Version]
  226. Juge, P.A.; Lee, J.S.; Ebstein, E.; Furukawa, H.; Dobrinskikh, E.; Gazal, S.; Kannengiesser, C.; Ottaviani, S.; Oka, S.; Tohma, S.; et al. MUC5B Promoter Variant and Rheumatoid Arthritis with Interstitial Lung Disease. N. Engl. J. Med. 2018, 379, 2209–2219. [Google Scholar] [CrossRef]
  227. Furusawa, H.; Peljto, A.L.; Walts, A.D.; Cardwell, J.; Molyneaux, P.L.; Lee, J.S.; Fernández Pérez, E.R.; Wolters, P.J.; Yang, I.V.; Schwartz, D.A. Common idiopathic pulmonary fibrosis risk variants are associated with hypersensitivity pneumonitis. Thorax 2022, 77, 508–510. [Google Scholar] [CrossRef]
  228. Okamoto, T.; Dobrinskikh, E.; Hennessy, C.E.; Liu, N.; Schwarz, M.I.; Evans, C.M.; Fontenot, A.P.; Yang, I.V.; Schwartz, D.A. Muc5b plays a role in the development of inflammation and fibrosis in hypersensitivity pneumonitis induced by Saccharopolyspora rectivirgula. Am. J. Physiol. Lung Cell Mol. Physiol. 2022, 323, L329–L337. [Google Scholar] [CrossRef]
  229. Ley, B.; Newton, C.A.; Arnould, I.; Elicker, B.M.; Henry, T.S.; Vittinghoff, E.; Golden, J.A.; Jones, K.D.; Batra, K.; Torrealba, J.; et al. The MUC5B promoter polymorphism and telomere length in patients with chronic hypersensitivity pneumonitis: An observational cohort-control study. Lancet Respir. Med. 2017, 5, 639–647. [Google Scholar] [CrossRef]
  230. Hobbs, B.D.; Putman, R.K.; Araki, T.; Nishino, M.; Gudmundsson, G.; Gudnason, V.; Eiriksdottir, G.; Zilhao Nogueira, N.R.; Dupuis, J.; Xu, H.; et al. Overlap of Genetic Risk between Interstitial Lung Abnormalities and Idiopathic Pulmonary Fibrosis. Am. J. Respir. Crit. Care Med. 2019, 200, 1402–1413. [Google Scholar] [CrossRef]
  231. Mathai, S.K.; Humphries, S.; Kropski, J.A.; Blackwell, T.S.; Powers, J.; Walts, A.D.; Markin, C.; Woodward, J.; Chung, J.H.; Brown, K.K.; et al. MUC5B variant is associated with visually and quantitatively detected preclinical pulmonary fibrosis. Thorax 2019, 74, 1131–1139. [Google Scholar] [CrossRef]
  232. Putman, R.K.; Gudmundsson, G.; Araki, T.; Nishino, M.; Sigurdsson, S.; Gudmundsson, E.F.; Eiríksdottír, G.; Aspelund, T.; Ross, J.C.; San José Estépar, R.; et al. The MUC5B promoter polymorphism is associated with specific interstitial lung abnormality subtypes. Eur. Respir. J. 2017, 50, 1700537. [Google Scholar] [CrossRef]
  233. Stancil, I.T.; Michalski, J.E.; Davis-Hall, D.; Chu, H.W.; Park, J.A.; Magin, C.M.; Yang, I.V.; Smith, B.J.; Dobrinskikh, E.; Schwartz, D.A. Pulmonary fibrosis distal airway epithelia are dynamically and structurally dysfunctional. Nat. Commun. 2021, 12, 4566. [Google Scholar] [CrossRef]
  234. Gally, F.; Sasse, S.K.; Kurche, J.S.; Gruca, M.A.; Cardwell, J.H.; Okamoto, T.; Chu, H.W.; Hou, X.; Poirion, O.B.; Buchanan, J.; et al. The MUC5B-associated variant rs35705950 resides within an enhancer subject to lineage- and disease-dependent epigenetic remodeling. JCI Insight 2021, 6, e144294. [Google Scholar] [CrossRef]
  235. Hancock, L.A.; Hennessy, C.E.; Solomon, G.M.; Dobrinskikh, E.; Estrella, A.; Hara, N.; Hill, D.B.; Kissner, W.J.; Markovetz, M.R.; Grove Villalon, D.E.; et al. Muc5b overexpression causes mucociliary dysfunction and enhances lung fibrosis in mice. Nat. Commun. 2018, 9, 5363. [Google Scholar] [CrossRef] [Green Version]
  236. Schwartz, D.A. Idiopathic Pulmonary Fibrosis Is a Genetic Disease Involving Mucus and the Peripheral Airways. Ann. Am. Thorac. Soc. 2018, 15, S192–S197. [Google Scholar] [CrossRef]
  237. Helling, B.A.; Gerber, A.N.; Kadiyala, V.; Sasse, S.K.; Pedersen, B.S.; Sparks, L.; Nakano, Y.; Okamoto, T.; Evans, C.M.; Yang, I.V.; et al. Regulation of MUC5B Expression in Idiopathic Pulmonary Fibrosis. Am. J. Respir. Cell Mol. Biol. 2017, 57, 91–99. [Google Scholar] [CrossRef]
  238. Molyneaux, P.L.; Cox, M.J.; Willis-Owen, S.A.; Mallia, P.; Russell, K.E.; Russell, A.M.; Murphy, E.; Johnston, S.L.; Schwartz, D.A.; Wells, A.U.; et al. The role of bacteria in the pathogenesis and progression of idiopathic pulmonary fibrosis. Am. J. Respir. Crit. Care Med. 2014, 190, 906–913. [Google Scholar] [CrossRef] [Green Version]
  239. Schwartz, D.A.; Blumhagen, R.Z.; Fingerlin, T.E. Evolution of the Gain-of-Function MUC5B Promoter Variant. Am. J. Respir. Crit. Care Med. 2022, 206, 1189–1191. [Google Scholar] [CrossRef]
  240. Borie, R.; Cardwell, J.; Konigsberg, I.R.; Moore, C.M.; Zhang, W.; Sasse, S.K.; Gally, F.; Dobrinskikh, E.; Walts, A.; Powers, J.; et al. Colocalization of Gene Expression and DNA Methylation with Genetic Risk Variants Supports Functional Roles of MUC5B and DSP in Idiopathic Pulmonary Fibrosis. Am. J. Respir. Crit. Care Med. 2022, 206, 1259–1270. [Google Scholar] [CrossRef]
  241. Kim, E.; Mathai, S.K.; Stancil, I.T.; Ma, X.; Hernandez-Gutierrez, A.; Becerra, J.N.; Marrero-Torres, E.; Hennessy, C.E.; Hatakka, K.; Wartchow, E.P.; et al. Aberrant Multiciliogenesis in Idiopathic Pulmonary Fibrosis. Am. J. Respir. Cell Mol. Biol. 2022, 67, 188–200. [Google Scholar] [CrossRef] [PubMed]
  242. Mota, P.C.; Soares, M.L.; Vasconcelos, C.D.; Ferreira, A.C.; Lima, B.A.; Manduchi, E.; Moore, J.H.; Melo, N.; Novais-Bastos, H.; Pereira, J.M.; et al. Predictive value of common genetic variants in idiopathic pulmonary fibrosis survival. J. Mol. Med. 2022, 100, 1341–1353. [Google Scholar] [CrossRef] [PubMed]
  243. Chung, J.H.; Peljto, A.L.; Chawla, A.; Talbert, J.L.; McKean, D.F.; Rho, B.H.; Fingerlin, T.E.; Schwarz, M.I.; Schwartz, D.A.; Lynch, D.A. CT Imaging Phenotypes of Pulmonary Fibrosis in the MUC5B Promoter Site Polymorphism. Chest 2016, 149, 1215–1222. [Google Scholar] [CrossRef] [Green Version]
  244. Chung, J.H.; Chawla, A.; Peljto, A.L.; Cool, C.D.; Groshong, S.D.; Talbert, J.L.; McKean, D.F.; Brown, K.K.; Fingerlin, T.E.; Schwarz, M.I.; et al. CT scan findings of probable usual interstitial pneumonitis have a high predictive value for histologic usual interstitial pneumonitis. Chest 2015, 147, 450–459. [Google Scholar] [CrossRef] [Green Version]
  245. Peljto, A.L.; Zhang, Y.; Fingerlin, T.E.; Ma, S.F.; Garcia, J.G.; Richards, T.J.; Silveira, L.J.; Lindell, K.O.; Steele, M.P.; Loyd, J.E.; et al. Association between the MUC5B promoter polymorphism and survival in patients with idiopathic pulmonary fibrosis. JAMA 2013, 309, 2232–2239. [Google Scholar] [CrossRef]
  246. van der Vis, J.J.; Snetselaar, R.; Kazemier, K.M.; ten Klooster, L.; Grutters, J.C.; van Moorsel, C.H. Effect of Muc5b promoter polymorphism on disease predisposition and survival in idiopathic interstitial pneumonias. Respirology 2016, 21, 712–717. [Google Scholar] [CrossRef]
  247. Hatabu, H.; Hunninghake, G.M.; Richeldi, L.; Brown, K.K.; Wells, A.U.; Remy-Jardin, M.; Verschakelen, J.; Nicholson, A.G.; Beasley, M.B.; Christiani, D.C.; et al. Interstitial lung abnormalities detected incidentally on CT: A Position Paper from the Fleischner Society. Lancet Respir. Med. 2020, 8, 726–737. [Google Scholar] [CrossRef]
  248. Araki, T.; Putman, R.K.; Hatabu, H.; Gao, W.; Dupuis, J.; Latourelle, J.C.; Nishino, M.; Zazueta, O.E.; Kurugol, S.; Ross, J.C.; et al. Development and Progression of Interstitial Lung Abnormalities in the Framingham Heart Study. Am. J. Respir. Crit. Care Med. 2016, 194, 1514–1522. [Google Scholar] [CrossRef] [Green Version]
  249. Juge, P.A.; Granger, B.; Debray, M.P.; Ebstein, E.; Louis-Sidney, F.; Kedra, J.; Doyle, T.J.; Borie, R.; Constantin, A.; Combe, B.; et al. A Risk Score to Detect Subclinical Rheumatoid Arthritis-Associated Interstitial Lung Disease. Arthritis Rheumatol. 2022, 74, 1755–1765. [Google Scholar] [CrossRef]
  250. Bonella, F.; Campo, I.; Zorzetto, M.; Boerner, E.; Ohshimo, S.; Theegarten, D.; Taube, C.; Costabel, U. Potential clinical utility of MUC5B und TOLLIP single nucleotide polymorphisms (SNPs) in the management of patients with IPF. Orphanet J. Rare Dis. 2021, 16, 111. [Google Scholar] [CrossRef]
  251. Oldham, J.M.; Ma, S.F.; Martinez, F.J.; Anstrom, K.J.; Raghu, G.; Schwartz, D.A.; Valenzi, E.; Witt, L.; Lee, C.; Vij, R.; et al. TOLLIP, MUC5B, and the Response to N-Acetylcysteine among Individuals with Idiopathic Pulmonary Fibrosis. Am. J. Respir. Crit. Care Med. 2015, 192, 1475–1482. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  252. Bush, A.; Pabary, R. Pulmonary alveolarproteinosis in children. Breathe 2020, 16, 200001. [Google Scholar] [CrossRef] [PubMed]
  253. Has, C.; Spartà, G.; Kiritsi, D.; Weibel, L.; Moeller, A.; Vega-Warner, V.; Waters, A.; He, Y.; Anikster, Y.; Esser, P.; et al. Integrin α3 mutations with kidney, lung, and skin disease. N. Engl. J. Med. 2012, 366, 1508–1514. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  254. Carlens, J.; Johnson, K.T.; Bush, A.; Renz, D.; Hehr, U.; Laenger, F.; Hogg, C.; Wetzke, M.; Schwerk, N.; Rayment, J.H. Heterogenous Disease Course and Long-term Outcome of Children’s Interstitial Lung Disease Related to Filamin A Gene Variants. Ann. Am. Thorac. Soc. 2022. [Google Scholar] [CrossRef] [PubMed]
  255. Griese, M. Etiologic Classification of Diffuse Parenchymal (Interstitial) Lung Diseases. J. Clin. Med. 2022, 11, 1747. [Google Scholar] [CrossRef]
  256. Bush, A. Paediatric interstitial lung disease: Not just kid’s stuff. Eur. Respir. J. 2004, 24, 521–523. [Google Scholar] [CrossRef] [Green Version]
  257. Borie, R.; Kannengiesser, C.; Antoniou, K.; Bonella, F.; Crestani, B.; Fabre, A.; Froidure, A.; Galvin, L.; Griese, M.; Grutters, J.C.; et al. European Respiratory Society Statement on Familial Pulmonary Fibrosis. Eur. Respir. J. 2022; in press. [Google Scholar] [CrossRef]
  258. Koucký, V.; Pohunek, P.; Vašáková, M.; Bush, A. Transition of patients with interstitial lung disease from paediatric to adult care. ERJ Open Res. 2021, 7, 00964–2020. [Google Scholar] [CrossRef]
  259. Newton, C.A.; Oldham, J.M.; Applegate, C.; Carmichael, N.; Powell, K.; Dilling, D.; Schmidt, S.L.; Scholand, M.B.; Armanios, M.; Garcia, C.K.; et al. The Role of Genetic Testing in Pulmonary Fibrosis: A Perspective From the Pulmonary Fibrosis Foundation Genetic Testing Work Group. Chest 2022, 162, 394–405. [Google Scholar] [CrossRef]
  260. Rosas, I.O.; Ren, P.; Avila, N.A.; Chow, C.K.; Franks, T.J.; Travis, W.D.; McCoy, J.P., Jr.; May, R.M.; Wu, H.P.; Nguyen, D.M.; et al. Early interstitial lung disease in familial pulmonary fibrosis. Am. J. Respir. Crit. Care Med. 2007, 176, 698–705. [Google Scholar] [CrossRef] [Green Version]
  261. Fernandez, B.A.; Fox, G.; Bhatia, R.; Sala, E.; Noble, B.; Denic, N.; Fernandez, D.; Duguid, N.; Dohey, A.; Kamel, F.; et al. A Newfoundland cohort of familial and sporadic idiopathic pulmonary fibrosis patients: Clinical and genetic features. Respir. Res. 2012, 13, 64. [Google Scholar] [CrossRef] [Green Version]
  262. Richeldi, L.; Collard, H.R.; Jones, M.G. Idiopathic pulmonary fibrosis. Lancet 2017, 389, 1941–1952. [Google Scholar] [CrossRef] [PubMed]
  263. Cottin, V.; Cordier, J.F. Velcro crackles: The key for early diagnosis of idiopathic pulmonary fibrosis? Eur. Respir. J. 2012, 40, 519–521. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  264. Cottin, V.; Richeldi, L. Neglected evidence in idiopathic pulmonary fibrosis and the importance of early diagnosis and treatment. Eur. Respir. Rev. 2014, 23, 106–110. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  265. Spagnolo, P.; Ryerson, C.J.; Putman, R.; Oldham, J.; Salisbury, M.; Sverzellati, N.; Valenzuela, C.; Guler, S.; Jones, S.; Wijsenbeek, M.; et al. Early diagnosis of fibrotic interstitial lung disease: Challenges and opportunities. Lancet Respir. Med. 2021, 9, 1065–1076. [Google Scholar] [CrossRef]
Figure 1. Major discoveries in the history of genetics: 1859: Ch. Darwin: “On the Origin of Species”; 1866: Gr. Mendel: Experiments in Plant Hybridization”; 1869: Fr. Miescher: “nuclein” (DNA); 1919: Ph. Levene: DNA structure; 1926: HJ Müller: radiation and lethal mutations; 1944: O. Avery, C. MacLeod, M. McCarty inheritance through “hereditary units”; 1950: E. Chargaff: “the total number of purines in DNA is equal to the total number of pyrimidines”; 1950s: JA Clements: physical properties of surfactant; 1953: J. Watson and F. Crick: three-dimensional double helical structure of DNA molecule; 1959: ME Avery: surfactant deficiency in RDS in premature babies; 1959: Hermansky–Pudlak syndrome; 1966: MW. Nirenberg break the genetic code; 1972: W. Fiers: first sequence of a gene; 1977: F. Sanger: sequence DNA for the first time; 1983: K. Banks Mullis: polymerase chain reaction; 2001–2003: Francis Collins: Human Genome Project; 2009: EH. Blackburn, CW. Greider, JW. Szostak: Nobel Prize “for the discovery of how chromosomes are protected by telomeres and the enzyme telomerase”.
Figure 1. Major discoveries in the history of genetics: 1859: Ch. Darwin: “On the Origin of Species”; 1866: Gr. Mendel: Experiments in Plant Hybridization”; 1869: Fr. Miescher: “nuclein” (DNA); 1919: Ph. Levene: DNA structure; 1926: HJ Müller: radiation and lethal mutations; 1944: O. Avery, C. MacLeod, M. McCarty inheritance through “hereditary units”; 1950: E. Chargaff: “the total number of purines in DNA is equal to the total number of pyrimidines”; 1950s: JA Clements: physical properties of surfactant; 1953: J. Watson and F. Crick: three-dimensional double helical structure of DNA molecule; 1959: ME Avery: surfactant deficiency in RDS in premature babies; 1959: Hermansky–Pudlak syndrome; 1966: MW. Nirenberg break the genetic code; 1972: W. Fiers: first sequence of a gene; 1977: F. Sanger: sequence DNA for the first time; 1983: K. Banks Mullis: polymerase chain reaction; 2001–2003: Francis Collins: Human Genome Project; 2009: EH. Blackburn, CW. Greider, JW. Szostak: Nobel Prize “for the discovery of how chromosomes are protected by telomeres and the enzyme telomerase”.
Diagnostics 12 02928 g001
Figure 2. Discoveries in IPF genetics: 1950: IPF is reported in identical twin sisters; 1996–1998: First mutations in Hermansky–Pudlak syndrome (HPS) genes; 1997: Is there is fibrosis gene in pulmonary fibrosis? 2001: First SFTPC mutation in familiar IP; 2002: NKX2-1 mutations in brain-thyroid-lung syndrome; 2003: HPS-4 mutations; 2004: Bi-allelic ABCA3 mutations in NRD syndrome and chILD; 2005: Mutations in TERT in patients with bone-marrow failure; 2007: TERT and TERC mutations in familiar IP; 2009: SFTPA2 mutations FIP and adenocarcinoma families, 2010: ELMOD2 SNP polymorphism in IPF; 2011: A common SNP in the promoter region of the gene MUC5B: the strongest polygenic risk factor for sporadic IPF and familiar IP; 2011: IL8 SNP polymorphism in IPF; 2012: CDKN1A, 1L1RN, TP53 SNP polymorphisms in IPF; 2013: ATP11A, DPP9, DSP, FAM13A, HLA-DRB1, MAPT, MUC2, OBFC1, TERC, TERT, MDGA2, SPPL2C, TGFB1, TOLLIP, TLR3 SNP polymorphisms in IPF; 2014: DKC1 mutation in familiar IP; 2014: TMEM173 mutations in SAVI, 2015: RTEL1, PARN and TINF2 mutations in familiar IP; 2015: COPA mutations in COPA syndrome; 2015: IND, LRRC34 SNP polymorphisms in IPF; 2016 SFTPA1 mutations in FIP and/or adenocarcinoma families; 2016: NAF1 mutations in pulmonary fibrosis-emphysema; 2017: AKAP13 SNP polymorphism in IPF; 2018: HPS-2 gene mutations: fibrosing lung disease early in childhood; 2019: ZCCHC8 mutations in FIP; 2020: NOP10 and NHP2 mutations in FIP; 2022: POT1 mutation in FIP; 2022: KNL1, NPRL3, STMN3, RTEL1, and an intergenic variant in 10q25.1 in IPF; 2022: KIF15 mutations in IPF; 2022: RPA1 mutations in pulmonary fibrosis and STS.
Figure 2. Discoveries in IPF genetics: 1950: IPF is reported in identical twin sisters; 1996–1998: First mutations in Hermansky–Pudlak syndrome (HPS) genes; 1997: Is there is fibrosis gene in pulmonary fibrosis? 2001: First SFTPC mutation in familiar IP; 2002: NKX2-1 mutations in brain-thyroid-lung syndrome; 2003: HPS-4 mutations; 2004: Bi-allelic ABCA3 mutations in NRD syndrome and chILD; 2005: Mutations in TERT in patients with bone-marrow failure; 2007: TERT and TERC mutations in familiar IP; 2009: SFTPA2 mutations FIP and adenocarcinoma families, 2010: ELMOD2 SNP polymorphism in IPF; 2011: A common SNP in the promoter region of the gene MUC5B: the strongest polygenic risk factor for sporadic IPF and familiar IP; 2011: IL8 SNP polymorphism in IPF; 2012: CDKN1A, 1L1RN, TP53 SNP polymorphisms in IPF; 2013: ATP11A, DPP9, DSP, FAM13A, HLA-DRB1, MAPT, MUC2, OBFC1, TERC, TERT, MDGA2, SPPL2C, TGFB1, TOLLIP, TLR3 SNP polymorphisms in IPF; 2014: DKC1 mutation in familiar IP; 2014: TMEM173 mutations in SAVI, 2015: RTEL1, PARN and TINF2 mutations in familiar IP; 2015: COPA mutations in COPA syndrome; 2015: IND, LRRC34 SNP polymorphisms in IPF; 2016 SFTPA1 mutations in FIP and/or adenocarcinoma families; 2016: NAF1 mutations in pulmonary fibrosis-emphysema; 2017: AKAP13 SNP polymorphism in IPF; 2018: HPS-2 gene mutations: fibrosing lung disease early in childhood; 2019: ZCCHC8 mutations in FIP; 2020: NOP10 and NHP2 mutations in FIP; 2022: POT1 mutation in FIP; 2022: KNL1, NPRL3, STMN3, RTEL1, and an intergenic variant in 10q25.1 in IPF; 2022: KIF15 mutations in IPF; 2022: RPA1 mutations in pulmonary fibrosis and STS.
Diagnostics 12 02928 g002
Table 1. Syndromic forms of pulmonary fibrosis.
Table 1. Syndromic forms of pulmonary fibrosis.
Telomeropathy
Dyskeratosis Congenita
Hermansky–Pudlak SAVI COPA
Clinical and Laboratory Manifestations
Skin: hyperpigmentation, nail dystrophy, hyperhidrosis, hair lossOculocutaneous albinism, photophobia, strabismus, nystagmusNon-healing ulcers
on cheeks, nose,
fingers, toes, soles
Diffuse alveolar hemorrhage
Lung fibrosis
Oral: Leukoplakia, ulcerations, atrophic glossitis, periodontitis, dental carries, lichenoid lesions, hyperpigmentation, neoplasmsBleeding disordersFailure to thrive in
children
Renal disease lupus nephritis
Lung fibrosis
Liver cirrhosis
Lung fibrosis (HPS1, HPS4, and AP3B1 genes)Lung fibrosisArthritis
Bone-marrow failure, anemiaKidney’s diseaseAnemia, leukopenia, thrombocytosis,
T-cell lymphopenia
Cystic lung disease
Skull and CNS: microcephalia, cerebral hypoplasia, mental retardation, deafnessGranulomatous colitisMyositis
Hypogonadism Joint stiffness
Esophageal stenosis
Osteoporosis
Short stature
Skeletal malformations
Epiphora
Genetic Variants
Interferonopathies
17genes implicated: ACD, CTC1, DCLRE1B, DKC1, NHP2, NOP10, PARN, POT1, RPA1, RTEL1, STN1, TERC, TERT, TINF2, WRAP53, and ZCCHC810 genes implicated: AP3B1, HPS1, HPS3, HPS4, HPS5, HPS6, and less commonly, AP3D1, BLOC1S3, BLOC1S6, and DTNBP1 inherited in an autosomal recessive mannerSTING1COPA
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Papiris, S.A.; Kannengiesser, C.; Borie, R.; Kolilekas, L.; Kallieri, M.; Apollonatou, V.; Ba, I.; Nathan, N.; Bush, A.; Griese, M.; et al. Genetics in Idiopathic Pulmonary Fibrosis: A Clinical Perspective. Diagnostics 2022, 12, 2928. https://doi.org/10.3390/diagnostics12122928

AMA Style

Papiris SA, Kannengiesser C, Borie R, Kolilekas L, Kallieri M, Apollonatou V, Ba I, Nathan N, Bush A, Griese M, et al. Genetics in Idiopathic Pulmonary Fibrosis: A Clinical Perspective. Diagnostics. 2022; 12(12):2928. https://doi.org/10.3390/diagnostics12122928

Chicago/Turabian Style

Papiris, Spyros A., Caroline Kannengiesser, Raphael Borie, Lykourgos Kolilekas, Maria Kallieri, Vasiliki Apollonatou, Ibrahima Ba, Nadia Nathan, Andrew Bush, Matthias Griese, and et al. 2022. "Genetics in Idiopathic Pulmonary Fibrosis: A Clinical Perspective" Diagnostics 12, no. 12: 2928. https://doi.org/10.3390/diagnostics12122928

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop